You are on page 1of 322

Renewable Energy

and Climate Change


Renewable Energy
and Climate Change

Volker Quaschning
Berlin University of Applied Systems HTW, Germany

(Translator) Hedy Jourdan

A John Wiley & Sons, Ltd., Publication


First published under the title Erneuerbare Energien und Klimaschutz by Carl Hanser
Verlag © 2008 Carl Hanser Verlag München
All Rights reserved.
Authorized translation from the 1st edition in the original German language published by
Carl Hanser Verlag München.
This edition first published 2010
© 2010 John Wiley & Sons, Ltd
Registered office
John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester, West Sussex, PO19 8SQ,
United Kingdom
For details of our global editorial offices, for customer services and for information about
how to apply for permission to reuse the copyright material in this book please see our
website at www.wiley.com.
The right of the author to be identified as the author of this work has been asserted in
accordance with the Copyright, Designs and Patents Act 1988.
All rights reserved. No part of this publication may be reproduced, stored in a retrieval
system, or transmitted, in any form or by any means, electronic, mechanical, photocopying,
recording or otherwise, except as permitted by the UK Copyright, Designs and Patents Act
1988, without the prior permission of the publisher.
Wiley also publishes its books in a variety of electronic formats. Some content that
appears in print may not be available in electronic books.
Designations used by companies to distinguish their products are often claimed as
trademarks. All brand names and product names used in this book are trade names, service
marks, trademarks or registered trademarks of their respective owners. The publisher is not
associated with any product or vendor mentioned in this book. This publication is designed
to provide accurate and authoritative information in regard to the subject matter covered. It
is sold on the understanding that the publisher is not engaged in rendering professional
services. If professional advice or other expert assistance is required, the services of a
competent professional should be sought.
Library of Congress Cataloging-in-Publication Data
Quaschning, Volker, 1969–
[Erneuerbare Energien und Klimaschutz. English]
Renewable energy and climate change / Volker Quaschning.
p. cm.
Includes bibliographical references and index.
ISBN 978-0-470-74707-0 (cloth : alk. paper) 1. Renewable energy sources.
2. Climatic changes. I. Title.
TJ808.Q38913 2010
363.738′74–dc22
2009045463
A catalogue record for this book is available from the British Library.
ISBN 978-0-470-74707-0
Set in 10 on 12 pt Times NR by Toppan Best-set Premedia Limited
Printed in Singapore by Fabulous Printers Pte Ltd
Contents

Preface........................................................................................................xi

1 Our Hunger for Energy ....................................................................1


1.1 Energy Supply – Yesterday and Today .......................................................... 2
1.1.1 From the French Revolution to the Early 20th Century ................ 2
1.1.2 The Era of Black Gold .................................................................... 4
1.1.3 Natural Gas – the Newest Fossil Energy Source ............................ 7
1.1.4 Nuclear Power – Split Energy ......................................................... 9
1.1.5 The Century of Fossil Energy ....................................................... 12
1.2 Energy Needs – Who Needs What, Where and How Much? ..................... 13
1.3 ‘Anyway’ Energy.......................................................................................... 16
1.4 Energy Supplies – Wealth Forever .............................................................. 19
1.5 The End of Fission ...................................................................................... 21
1.6 Oil Prices Today – Politics, Supply and Demand ....................................... 22

2 The Climate Before the Collapse? ...............................................24


2.1 It Is Getting Warm – Climate Changes Today ............................................ 24
2.1.1 The Ice is Slowly Melting ............................................................. 24
2.1.2 Natural Catastrophes Occur More Frequently .............................. 26
2.2 The Guilty Parties – Causes of Climate Change ........................................ 29
2.2.1 The Greenhouse Effect .................................................................. 29
2.2.2 The Prime Suspect: Carbon Dioxide ............................................. 30
2.2.3 Other Culprits ................................................................................ 34
2.3 Outlook and Recommendations – What Lies Ahead? ................................ 37
2.3.1 Will It be Bitterly Cold in Europe? ............................................... 39
2.3.2 Recommendations for Effective Climate Protection ..................... 42
2.4 Difficult Birth – Politics and Climate Change ............................................ 42
2.4.1 German Climate Policy.................................................................. 42
2.4.2 International Climate Policy .......................................................... 44
2.5 Self-Help Climate Protection ...................................................................... 46

v
Contents

3 From Wasting Energy to Saving Energy and Reducing


Carbon Dioxide ...............................................................................47
3.1 Less Efficient – Energy Use and Waste Today ........................................... 47
3.2 Personal Energy Needs – Easily Saved at Home ........................................ 50
3.2.1 Domestic Electricity – Money Wasted .......................................... 50
3.2.2 Heat – Surviving the Winter with Almost No Heating ................. 54
3.2.3 Transport – Getting Somewhere Using Less Energy .................... 58
3.3 Industry and Co – Everyone Else is to Blame ........................................... 61
3.4 The Personal Carbon Dioxide Record ........................................................ 62
3.4.1 Emissions Caused Directly by One’s Own Activities ................... 62
3.4.2 Indirectly Caused Emissions ......................................................... 63
3.4.3 Total Emissions .............................................................................. 65
3.5 The Sale of Ecological Indulgences ............................................................ 67

4 Carbon-Free Energy – Vision or Utopia? ....................................70


4.1 Options for Carbon-Free Energy Supply..................................................... 70
4.1.1 Efficient Power Plants – More Power with Less
Carbon Dioxide .............................................................................. 70
4.1.2 Carbon Dioxide Sequestering – Away with
Carbon Dioxide .............................................................................. 72
4.1.3 Nuclear Energy – Squeaky Clean.................................................. 74
4.1.4 Combined Heat and Power – Using Fuel Twice ........................... 75
4.1.5 Saving Energy – Achieving More with Less ................................ 76
4.2 Renewable Energy Sources – No End to What is Available ....................... 77
4.3 Options for Protecting the Climate ............................................................. 79
4.3.1 Down with Primary Energy Needs................................................ 79
4.3.2 Electricity Generation Totally Without Nuclear and Fossil
Power Plants ................................................................................... 81
4.3.3 Insulation and Renewable Energies to Provide Heat .................... 82
4.3.4 Increasing Efficiency and New Concepts for Traffic .................... 83
4.4 Reliable Supply Using Renewable Energies ............................................... 84

5 Photovoltaics – Energy from Sand ..............................................87


5.1 Structure and Function ................................................................................ 88
5.1.1 Electrons, Holes and Space-Charge Regions ................................ 88
5.1.2 Efficiency, Characteristics and MPP ............................................. 90
5.2 Production of Solar Cells – from Sand to Cell ........................................... 92
5.2.1 Silicon Solar Cells – Electricity from Sand .................................. 92
5.2.2 From Cell to Module ..................................................................... 94
5.2.3 Thin Film Solar Cells .................................................................... 95
5.3 Photovoltaic Systems – Networks and Islands ............................................ 96
5.3.1 Sun Islands ..................................................................................... 96
5.3.2 Sun in the Grid .............................................................................. 99
5.4 Planning and Design .................................................................................. 103
5.4.1 Planned on the Grid ..................................................................... 103
5.4.2 Planned Islands ............................................................................ 107

vi
Contents

5.5 Economics .................................................................................................. 109


5.5.1 What Does It Cost?...................................................................... 109
5.5.2 Incentive Schemes ....................................................................... 111
5.6 Ecology ...................................................................................................... 112
5.7 Photovoltaic Markets ................................................................................. 113
5.8 Outlook and Development Potential.......................................................... 114

6 Solar Thermal Systems – Year-Round Heating from the Sun ... 116
6.1 Structure and Functionality ....................................................................... 118
6.2 Solar Collectors – Collecting the Sun ....................................................... 120
6.2.1 Swimming Pool Absorbers ......................................................... 120
6.2.2 Flat-Plate Collectors .................................................................... 121
6.2.3 Air-Based Collectors.................................................................... 122
6.2.4 Vacuum-Tube Collectors.............................................................. 123
6.3 Solar Thermal Systems ............................................................................. 125
6.3.1 Hot Water from the Sun............................................................... 125
6.3.2 Heating with the Sun ................................................................... 128
6.3.3 Solar Communities ...................................................................... 130
6.3.4 Cooling with the Sun .................................................................. 130
6.3.5 Swimming with the Sun .............................................................. 131
6.3.6 Cooking with the Sun .................................................................. 133
6.4 Planning and Design ................................................................................. 133
6.4.1 Solar Thermal Heating of Domestic Hot Water .......................... 134
6.4.2 Solar Thermal Heating as Support Heating ................................ 136
6.5 Economics ................................................................................................. 138
6.6 Ecology ...................................................................................................... 139
6.7 Solar Thermal Markets ............................................................................. 140
6.8 Outlook and Development Potential.......................................................... 142

7 Solar Power Plants – Even More Energy from the Sun ...........144
7.1 Concentration on the Sun .......................................................................... 145
7.2 Solar Power Plants ..................................................................................... 147
7.2.1 Parabolic Trough Power Plants .................................................... 147
7.2.2 Solar Tower Power Plants ............................................................ 150
7.2.3 Dish-Stirling Power Plants ........................................................... 153
7.2.4 Solar Chimney Power Plants ....................................................... 153
7.2.5 Concentrating Photovoltaic Power Plants .................................... 155
7.2.6 Solar Chemistry ........................................................................... 155
7.3 Planning and Design .................................................................................. 156
7.3.1 Concentrating Solar Thermal Power Plants ................................. 157
7.3.2 Solar Chimney Power Plants ....................................................... 158
7.3.3 Concentrating Photovoltaic Power Plants .................................... 158
7.4 Economics .................................................................................................. 158
7.5 Ecology ...................................................................................................... 160
7.6 Solar Power Plant Markets ........................................................................ 161
7.7 Outlook and Development Potential.......................................................... 162

vii
Contents

8 Wind Power Systems – Electricity from Thin Air .....................165


8.1 Gone with the Wind – Where the Wind Comes From .............................. 166
8.2 Utilizing Wind............................................................................................ 168
8.3 Installations and Parks ............................................................................... 173
8.3.1 Wind Chargers ............................................................................. 173
8.3.2 Grid-Connected Wind Turbines ................................................... 174
8.3.3 Wind Farms .................................................................................. 178
8.3.4 Offshore Wind Farms................................................................... 179
8.4 Planning and Design .................................................................................. 182
8.5 Economics .................................................................................................. 184
8.6 Ecology ...................................................................................................... 187
8.7 Wind Power Markets.................................................................................. 188
8.8 Outlook and Development Potential.......................................................... 189

9 Hydropower Plants – Wet Energy ..............................................191


9.1 Tapping into the Water Cycle .................................................................... 192
9.2 Water Turbines ........................................................................................... 194
9.3 Hydropower Plants..................................................................................... 197
9.3.1 Run-of-River Power Plants .......................................................... 197
9.3.2 Storage Power Plants ................................................................... 198
9.3.3 Pumped-Storage Power Plants ..................................................... 200
9.3.4 Tidal Power Plants ....................................................................... 201
9.3.5 Wave Power Plants ....................................................................... 202
9.3.6 Ocean Current Power Plants ........................................................ 203
9.4 Planning and Design .................................................................................. 204
9.5 Economics .................................................................................................. 206
9.6 Ecology ...................................................................................................... 206
9.7 Hydropower Markets ................................................................................. 207
9.8 Outlook and Development Potential.......................................................... 209

10 Geothermal Energy – Power from the Deep .............................210


10.1 Tapping into the Earth’s Heat .................................................................... 210
10.2 Geothermal Heat and Power Plants ........................................................... 215
10.2.1 Geothermal Heat Plants ............................................................... 215
10.2.2 Geothermal Power Plants............................................................. 216
10.2.3 Geothermal HDR Power Plants ................................................... 218
10.3 Planning and Design .................................................................................. 219
10.4 Economics .................................................................................................. 220
10.5 Ecology ...................................................................................................... 220
10.6 Geothermal Markets .................................................................................. 221
10.7 Outlook and Development Potential.......................................................... 222

11 Heat Pumps – from Cold to Hot .................................................223


11.1 Heat Sources for Low-Temperature Heat .................................................. 223
11.2 Working Principle of Heat Pumps ............................................................. 226

viii
Contents

11.2.1 Compression Heat Pumps............................................................ 226


11.2.2 Absorption Heat Pumps and Adsorption Heat Pumps................ 227
11.3 Planning and Design .................................................................................. 228
11.4 Economics .................................................................................................. 232
11.5 Ecology ...................................................................................................... 233
11.6 Heat Pump Markets ................................................................................... 235
11.7 Outlook and Development Potential.......................................................... 236

12 Biomass – Energy from Nature ..................................................237


12.1 Origins and Use of Biomass...................................................................... 238
12.2 Biomass Heating ........................................................................................ 241
12.2.1 Wood as a Fuel ............................................................................ 241
12.2.2 Fireplaces and Closed Woodburning Stoves ............................... 245
12.2.3 Firewood Boilers .......................................................................... 245
12.2.4 Wood Pellet Heating .................................................................... 246
12.3 Biomass Heat and Power Plants ................................................................ 248
12.4 Biofuels ...................................................................................................... 250
12.4.1 Bio-oil .......................................................................................... 251
12.4.2 Biodiesel ...................................................................................... 251
12.4.3 Bioethanol .................................................................................... 252
12.4.4 BtL Fuels ..................................................................................... 253
12.4.5 Biogas........................................................................................... 254
12.5 Planning and Design .................................................................................. 255
12.5.1 Firewood Boilers .......................................................................... 256
12.5.2 Wood Pellet Heating .................................................................... 256
12.6 Economics .................................................................................................. 259
12.7 Ecology ...................................................................................................... 260
12.7.1 Solid Fuels ................................................................................... 260
12.7.2 Biofuels ........................................................................................ 262
12.8 Biomass Markets ....................................................................................... 263
12.9 Outlook and Development Potential.......................................................... 264

13 The Hydrogen Industry and Fuel Cells ......................................265


13.1 Hydrogen as an Energy Source ................................................................. 266
13.1.1 Production of Hydrogen .............................................................. 267
13.1.2 Storage and Transport of Hydrogen ............................................ 269
13.2 Fuel Cells: Bearers of Hope ...................................................................... 270
13.3 Economics .................................................................................................. 272
13.4 Ecology ...................................................................................................... 273
13.5 Markets, Outlook and Development Potential .......................................... 274

14 Sunny Prospects – Examples of Sustainable


Energy Supply ..............................................................................276
14.1 Climate-Compatible Living ....................................................................... 276
14.1.1 Carbon-Neutral Standard Prefabricated Houses ......................... 277

ix
Contents

14.1.2 Plus-Energy Solar House ............................................................. 278


14.1.3 Plus-Energy Housing Estate ........................................................ 279
14.1.4 Heating Only with the Sun .......................................................... 279
14.1.5 Zero Heating Costs after Redevelopment ................................... 280
14.2 Working and Producing in Compatibility with the Climate ..................... 281
14.2.1 Offices and Shops in Solar Ship.................................................. 281
14.2.2 Zero-Emissions Factory ............................................................... 282
14.2.3 Carbon-Free Heavy Equipment Factory ...................................... 283
14.3 Climate-Compatible Driving ..................................................................... 284
14.3.1 Waste Gas-Free Electropower...................................................... 284
14.3.2 Travelling around the World in a Solar Mobile .......................... 285
14.3.3 Across Australia in Thirty-Three Hours ...................................... 286
14.3.4 Game over CO2!........................................................................... 287
14.4 Climate-Compatible Travel by Water or Air ............................................. 288
14.4.1 Modern Shipping ......................................................................... 288
14.4.2 Solar Ferry on Lake Constance ................................................... 289
14.4.3 World Altitude Record with a Solar Aeroplane .......................... 290
14.4.4 Flying around the World in a Solar Plane ................................... 291
14.4.5 Flying for Solar Kitchens ............................................................ 292
14.5 Carbon-Free Electricity for an Island ........................................................ 293
14.6 All’s Well that Ends Well ........................................................................... 294

Appendix .................................................................................................296
A.1 Energy Units and Prefixes ......................................................................... 296
A.2 Geographic Coordinates of Energy Power Plants ..................................... 297

References ..............................................................................................300

Index ........................................................................................................303

x
Preface

The problems of energy and climate change have finally ended up where they
belong: at the heart of public attention. Yet the connection between energy use and
global warming is something we have been aware of for decades. In the late 1980s
the German federal government proclaimed climate protection to be one of its main
targets. At the time numerous experts were already calling for a speedy restructuring
of the entire energy supply. Despite the government’s declaration, the official
response was, at best, half-hearted. But the climate problem can no longer remain
on the back burner. There is a growing awareness that climate change has already
begun. The prognosis of the researchers studying what is happening to our climate
is horrendous. If we do not pull the emergency cord soon, the catastrophic conse-
quences of climate change will far exceed even our powers of imagination. The
awarding of the Nobel Peace Prize to Al Gore, the US climate activist, and the
Intergovernmental Panel on Climate Change, which has been urgently warning of
the consequences for years, could be seen as a sign of helplessness rather than
optimism about solving the problem.
At the same time as climate change is threatening our environment, new records for
rising oil and natural gas prices show that the supplies still available will not be
enough to cover our requirements for much longer and that other alternatives must
be exploited as soon as possible.
And yet the solution is a simple one: renewable energy. Renewable energy could
completely cover all our energy supply needs within a few decades. This is the only
way to end our dependence on energy sources like oil and uranium, which are so
costly both in financial terms and in the havoc they wreak on our environment, and
satisfy our hunger for energy in a way that is sustainable and compatible with the
climate.
However, the path we need to take to get to that point is still unclear to many. Many
people still do not believe renewable energy offers a viable option. Some underes-
timate the alternative possibilities offered to such an extent that they predict a return
to the Stone Age once oil and coal supplies have been fully depleted.
The aim of this book is to eliminate these prejudices. It describes, clearly and simply,
the different technologies that exist and the potential for using renewable energy.

xi
Preface

The focus is always on the interaction between the different technologies. The
example of Germany shows the forms that sustainable energy supply can take and
how it can be implemented. But the book is designed to show all readers, wherever
they live, how they themselves can make a contribution towards building a climate-
compatible energy economy. In addition to explaining different energy measures
that individuals themselves can undertake, the book provides concrete planning aids
for implementing renewable energy systems.
This book has been specifically written so that it offers essential information to a
broad spectrum of readers. It introduces the different technologies to readers who
are new to the subject but at the same time provides interesting background infor-
mation to those who already have some knowledge about the field.
This book has been translated from the German version. It is an important supple-
ment to the technical book ‘Renewable Energy Systems’, written by me and pub-
lished by Hanser Verlag publishers. It is clear from the high level of interest
generated by this technical book, which is now in its fifth edition in German and
has been translated into English and Arabic, that a real need exists for literature on
the subject of renewable energy. The feedback I have received from the book and
from many of my lectures indicates that readers want something that offers an
overview of the subject that is easy to understand but still comprehensive. This book
should fill this gap and provide support in the development of sustainable energy
supply.

Berlin, 2009 Prof. Volker Quaschning


HTW Berlin – University of Applied Sciences
www.volker-quaschning.de

xii
1
1 Our Hunger for Energy
Most people have heard of the cult TV series Star Trek. Thanks to this programme,
we know that in the not-too-distant future man will start exploring the infinite
expanses of the universe. The energy issue will have been resolved long before
then. The Warp drive discovered in 2063 provides unlimited energy that Captain
Kirk uses to steer the starship Enterprise at speeds faster than light to new
adventures. Energy is available in overabundance; peace and prosperity rule on
earth and environmental problems are a thing of the past. But even this type of
energy supply is not totally without its risks. A warp core breach can cause as
much damage as a core meltdown in an ancient nuclear power station. Warp
plasma itself is not a totally safe material, as the regular viewers of Star Trek very
well know.
Unfortunately – or sometimes fortunately – most science fiction is far removed from
the real world. From our perspective the discovery of a warp drive seems highly
unlikely, even if dyed-in-the-wool Star Trek fans would like to think otherwise. We
are currently not even close to mastering comparatively simple nuclear fusion.
Consequently, we must rely on known technology, whatever its drawbacks, to solve
our energy problems.
In reality, energy use has always had a noticeable impact on the environment.
Looking back today, it is obvious that burning wood was less than ideal and that
the harmful noxious fumes created by such fires considerably reduced the life
expectancy of our ancestors. A fast-growing world population, increasing prosperity
and the hunger for fuel that has developed as a consequence have led to a rapid
rise in the need for energy. Although the resulting environmental problems may
only have affected certain regions, the effects of our hunger for energy can now
be felt around the world. Overconsumption of energy is the main trigger for the
global warming that is now threatening to cause devastation in many areas of the
world. However, resignation and fear are the wrong responses to this ever-growing
problem. There are alternative energy sources to be tapped. It is possible to develop
a long-term safe and affordable energy supply that will have only a minimal and

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

1
1 Our Hunger for Energy

manageable impact on the environment. This book describes the form this energy
supply must take and how each individual can contribute towards a collective effort
to halt climate change. But first it is important to take a close look at the causes of
today’s problems.

1.1 Energy Supply – Yesterday and Today

1.1.1 From the French Revolution to the Early 20th Century


At the time of the French Revolution in 1789 animal muscle power was the most
important source of energy. Around 14 million horses and 24 million cattle with an
overall output of around 7.5 billion watts were being used as work animals (König,
1999). This corresponds to the power of more than 100 000 mid-range cars.

Power and Energy or the Other Way Around


The terms ‘power’ and ‘energy’ are closely linked, and for this reason they
are often confused with one another and used incorrectly.
Energy is stored work; thus the possibility to perform work. It is identified by the symbol
E. The symbol for work is W.
Power (symbol: P) indicates the time during which the work is to be performed or the
energy used.

P=
W ⎛ power = work ⎞
⎜⎝ ⎟
t time ⎠
For example, if a person lifts a bucket of water, this is considered work. The work that
is performed increases the potential energy of the bucket of water. If the bucket is lifted
up twice as quickly, less time is used and the power is doubled, even if the work is
the same.
The unit for power is the watt (abbreviation: W). (The fact that the abbreviation for watt
is the same as the symbol for work does not simplify matters.)
The unit for energy is watt second (Ws) or joule (J). Other units are also used for energy.
Appendix A.1 provides the conversion factors between the different units of energy.
As the required powers and energies are often very high, prefixes such as mega (M),
giga (G), tera (T), peta (P) and exa (E) are frequently used (see Appendix A.1).

The second staple energy source in those days was firewood, which was so impor-
tant that it probably changed the political face of Europe. It is believed today that
the transfer of the Continent’s centre of power from the Mediterranean to north of
the Alps came about because of the abundance of forests and associated energy
potential there. Although the Islamic world was able to maintain its position of
power on the Iberian peninsula well into the 15th century, one of the reasons why
it lost its influence was the lack of wood. The problem was that there was not enough

2
1.1 Energy Supply – Yesterday and Today

firewood that could be used to melt down metal to produce cannons and other
weapons. This goes to show that energy crises are not just a modern phenomenon
(Figure 1.1).

Figure 1.1 Firewood, draught animals, wind and water power supplied most of the energy needed in
the world as late as the 18th century.

In addition to muscle power and firewood, other renewable energies were used
intensively until the beginning of the 20th century. Between 500 000 and 600 000
watermills were in operation in Europe at the end of the 18th century. The use of
wind power was also widespread, particularly in flat areas with a lot of wind. For
example, the United Netherlands had around 8000 working windmills at the end of
the 17th century.
For a long time fossil energy sources were only of secondary importance. Although
coal from underground deposits was known to be a source of energy, it was largely
avoided. It was not until a lack of wood in certain areas of Europe led to energy
shortages that coal deposits began to be exploited. In addition, the higher energy
density of coal proved to be an advantage in the production of steel. There was no
stopping the development of this resource once the industrial revolution dawned. In
1800 60% of coal was used to provide domestic heat, but 40 years later far more
coal was used in ironworks and other factories than in homes.

3
1 Our Hunger for Energy

Fossil Energy Sources – Stored Solar Energy


Fossil energy sources are concentrated energy sources that evolved from
animal and plant remains over very long periods of time. These sources
include oil, gas, hard coal, brown coal and turf. The base materials for fossil energy
sources could only develop because of their conversion through solar radiation over
millions of years. In this sense, fossil energy sources are a form of stored solar energy.
From a chemical point of view, fossil energy sources are based on organic carbon com-
pounds. Burnt in conjunction with oxygen, they not only generate energy in the form of
heat but also always produce the greenhouse gas carbon dioxide as well as other exhaust
gases.

In around 1530, coal mines in Great Britain were producing about 200 000 tons of
coal annually. By 1750 it was about 5 million tons and in 1854 an astonishing 64
million tons. By 1900 three countries, Britain, the USA and Germany, had an 80%
share of world production (König, 1999).

Renewable Energies – Not That New


The supplies of fossil energies, such as oil, natural gas and coal, are limited.
They will be depleted within a few decades and then cease to exist. Renewable
energy sources, on the other hand, ‘renew’ themselves on their own. For example, if a
hydropower plant takes the power of the water from a river, the river will not stop
flowing. The energy content of the river renews itself on its own because the sun evapo-
rates the water and the rain feeds the river again.
Renewable energies are also referred to as ‘regenerative’ or ‘alternative’ energies. Other
renewable energies include wind power, biomass, the natural heat of the earth and solar
energy. Even the sun will eventually disappear in around four billion years. Compared
to the few decades that fossil energy sources will still be available to us, this time period
seems infinitely long.
Incidentally, renewable energies have been used by mankind for considerably longer
than fossil fuels, although the current systems for using these fuels are vastly more
advanced than in the past. Therefore, it is not renewable energies that are new but rather
the knowledge that in the long term renewable energies are the only option for a safe
and environmentally compatible energy supply.

At the end of the 20th century worldwide coal production reached almost 4 billion
tons. With an overall share of less than 3% of the world market, Germany and
Britain had lost their former position of supremacy in the coal industry. Today
power stations use most of the available coal. China and the USA are currently the
main coal-producing countries by a considerable margin.

1.1.2 The Era of Black Gold


Like coal, oil consists of conversion products from animal and plant substances, the
biomass of primeval times. Over millions of years plankton and other single-celled

4
1.1 Energy Supply – Yesterday and Today

organisms were deposited in sea basins. Due to the lack of oxygen, they were unable
to decompose. Chemical processes of transformation eventually turned these sub-
stances into oil and gas. On the other hand, the biomass that was originally deposited
originated from the sun, which means that fossil energy sources like coal, oil and
gas are nothing more than long-term conservers of solar energy. The oldest oil
deposits are around 350 million years old. The area around the Persian Gulf where
most oil is exploited today was completely below sea level ten to fifteen million
years ago.
The oil deposits were developed much later than coal, because for a long time there
were no practical uses for the liquid energy source. Oil was used in small quantities
for thousands of years for medicinal and lighting purposes, but its high flammability
compared to coal and charcoal gave it the reputation of being a very dangerous fuel.
Petroleum lamps and later the invention of internal combustion engines finally
provided a breakthrough at the end of the 19th century.
Industrial oil production began in August 1859. While drilling at a depth of 20 m
near Titusville in the US state of Pennsylvania, the American Edwin L. Drake struck
oil. One name in particular is linked with further oil exploitation in America: John
Davison Rockefeller. In 1862 at the age of 23 he founded an oil company that
became Standard Oil and later the Exxon Corporation and incorporated large sec-
tions of the American oil industry.
However, it was still well into the 20th century before fossil energy supplies, and
specifically oil, dominated the energy market. In 1860 about 100 000 tons of oil
were produced worldwide; by 1895 it was already 14 million tons. German govern-
ment figures reveal that in 1895 there were 18 362 wind engines, 54 529 water
engines, 58 530 steam engines and 21 350 internal combustion engines in use in the
country (Gasch and Twele, 2004). Half the drive units even then were themselves
still being operated using renewable energy sources.
There was a huge rise in oil production in the 20th century. In 1929 output had
already risen to over 200 million tons and in the 1970s it shot up to over three billion
tons (Figure 1.2). Today oil is the most important energy source of most industrial-
ized countries. An average UK citizen, including small children and pensioners, uses
around 1600 litres of oil per year. An average German citizen uses 2000 litres and
an US citizen an astonishing 4000 litres. This amounts to 40 well-filled bathtubs.
Being too dependent on a single energy source can become a serious problem for
a society, as history shows. In 1960 OPEC (Organization of Petroleum Exporting
Countries) was founded with headquarters in Vienna. The goal of OPEC is to coor-
dinate and standardize the oil policies of its member states. These include Algeria,
Ecuador, Gabon, Indonesia, Iraq, Qatar, Kuwait, Libya, Nigeria, Saudi Arabia,
Venezuela and the United Arab Emirates, which at the end of the 20th century
together controlled 40% of oil production worldwide. As a result of the Yom Kippur
war between Israel, Syria and Egypt, the OPEC states cut back on production in
1973. This led to the first oil crisis and a drastic rise in oil prices. Triggered by
shortfalls in production and uncertainty after the revolution in Iran and the subse-
quent first Gulf war, the second oil crisis occurred in 1979 with oil prices rising to
US$38 per barrel.

5
1 Our Hunger for Energy

4000
oil crises
3500
global crude oil production in Mt

3000

2500

2000

1500
Great
Depression
1000

500

0
1860 1870 1880 1890 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000

Figure 1.2 Oil production since 1860 (Quaschning, 2007).

The drastic rise in oil prices set back world economic growth and energy use by
about four years. The industrialized nations, which had become used to low oil
prices, reacted with shock. Schemes such as car-free Sundays and programmes
promoting the use of renewable energies were the result. Differences between the
individual OPEC states in turn led to a rise in production quotas and a steep drop
in price at the end of the 1980s. This also sharply reduced the commitment of the
industrialized nations to use renewable energies.

From Alsatian Herring Barrels to Oil Barrels


Commercial oil production in Europe began in Pechelbronn in Alsace (now
France) in 1735. Cleaned herring casks were used to store the oil, because
in those days salted herring was traded in large quantities in barrels and so these barrels
were comparatively cheap. As oil production increased, special barrels of the same size
were produced exclusively for oil. The bottom of the barrels was painted blue to prevent
any confusion with barrels used for food products. When commercial oil production
began in the USA, the companies copied the techniques used in the Alsace region. This
also included the standard size of herring barrels. Since then the herring barrel volume
has remained the international measuring unit for oil. The abbreviation of barrel is bbl,
which stands for ‘blue barrel’ and means a barrel with a blue base.
1 petroleum barrel (US) = 1 bbl (US) = 158.987 l (litres )

The dramatic collapse in the price of crude oil from almost US$40 a barrel to US$10
created economic problems for some of the production countries and also made it
unattractive to develop new oil sources. In 1998 unity was largely restored again
among the OPEC states. They agreed on lower production quotas in order to halt
any further drop in prices. In fact, prices rose even higher than originally intended.

6
1.1 Energy Supply – Yesterday and Today

Now the lack of investment in energy-saving measures was coming home to roost.
The economic boom in China and in other countries further boosted the demand for
oil to such an extent that it was difficult to meet it. The consequence was that oil
prices kept climbing to new record highs. Even though the oil price has fallen
sharply again due to the current financial crisis, new record prices are expected again
due to the limited supplies available.
Yet there have been some fundamental changes since the beginning of the 1980s.
Energy use has decreased despite rapid and sustained economic growth. The realiza-
tion has set in that energy use and gross national product are not inextricably linked.
It is possible for prosperity to increase even if energy use stagnates or drops.
Nonetheless, the chance to develop true alternatives to oil and use energy-savings
options was missed due to the long period of continuous low oil prices. This is
particularly apparent in the transport sector. Cars became faster, more comfortable,
heavier and had more horsepower, but were only minimally more fuel-efficient. As
a result of all the talk about climate change and high oil prices, the car manufactur-
ers are now scrambling to incorporate features into their cars that have not been
demanded in decades: fuel efficiency and low emission of greenhouse gases.
As important as it is as a fuel, that is not the only use for oil, because it is also an
important raw material for the chemical industry. For example, oil is used as a basic
material in the production of plastic chairs, plastic bags, nylon tights, polyester
shirts, shower gels, scents and vitamin pills.

1.1.3 Natural Gas – the Newest Fossil Energy Source


Natural gas is considered to be the cleanest fossil energy source. When natural gas
is burnt, it produces fewer harmful substances and climate-damaging carbon dioxide
than oil or coal.
The base material for the creation of natural gas was usually green plants in the flat
coastal waters of the tropics. The Northern German lowland plains were part of this
area 300 million years ago. The lack of oxygen in coastal swampland prevented the
organic material from decomposing and so it developed into turf. As time went by,
new layers of sand and clay were deposited on the turf, which during the course of
millions of years turned into brown and bituminous coal. Natural gas then developed
from this due to the high pressure that exists at depths of several kilometres and
temperatures of 120 to 180 °C.
However, natural gas does not consist of a single gas, but rather a mixture of dif-
ferent gases whose composition varies considerably depending on the deposit. The
main component is methane, and the gas also often contains large quantities of
sulphur hydrogen, which is poisonous and even in very small concentrations smells
of rotten eggs. Therefore, natural gas must often first be purified in processing plants
using physiochemical processes. As natural gas deposits usually also contain water,
the gas must be dried to prevent corrosion in the natural gas pipelines.
Natural gas was not seen as a significant energy source until relatively recently. It
was not until the early 1960s that natural gas was promoted and marketed in large
quantities. The reasons for this late use of natural gas compared to coal and oil is

7
1 Our Hunger for Energy

Figure 1.3 Left: Building a natural gas pipeline in Eastern Germany. Right: Storage facility for 4.2
billion m3 of natural gas in Rehden, 60 km south of Bremen. Photos: WINGAS GmbH.

that extracting it requires drilling to depths of several thousand metres. It also


requires complicated transport. Whereas oil initially was still being transported in
wooden barrels, gas requires pressure storage or pipelines for its transport. Nowadays,
pipelines extend for thousands of kilometres from the places where the gas is
extracted all the way to where it provides gas heating to family homes (Figure 1.3).
The world’s largest gas producer is Russia, followed by the USA, Canada, Iran,
Norway and Algeria.
However, the demand for natural gas is not constant over the whole year. In coun-
tries with cold winters the demand in winter is often double what it is in summer.
As it is not economical to cut production by half in the summer, enormous storage
facilities are necessary to balance the uneven demand between summer and winter.
So-called salt caverns and aquifer reservoirs are used. Caverns are shafts dug in
underground salt deposits from where the stored gas can quickly be extracted – for
instance, to cover sudden high demand. Underground aquifer reservoirs are suitable
for the storage of particularly large quantities of gas. In total, Germany has over 30
billion cubic metres of natural gas in storage. Environmentally compatible hydrogen
is expected to play an important role in future energy supply in just a few decades’
time. It might be possible to convert the current natural gas storage facilities to store
the hydrogen.

8
1.1 Energy Supply – Yesterday and Today

1.1.4 Nuclear Power – Split Energy


In December 1938 Otto Hahn and Fritz Strassmann split a uranium nucleus on a
simple laboratory bench at the Kaiser-Wilhelm Institute for Chemistry in Berlin-
Dahlem, thereby laying the foundation for the future use of nuclear energy. The
laboratory bench can still be admired today at the Deutsches Museum in Munich.
In the experiment a uranium-235 nucleus was bombarded with slow neutrons. The
nucleus then split and produced two atomic parts, krypton and barium, as well as
two or three other neutrons. With a large quantity of uranium-235, these new neu-
trons can also split uranium nuclei that in turn release neutrons, thus leading to a
chain reaction. If enough uranium is available, the uncontrolled chain reaction will
create an atomic bomb. If the speed of the chain reaction can be controlled, uranium-
235 can also be used as fuel for power stations.
A so-called mass defect exists with nuclear fission. The mass of all the little pieces
after the fission is less than that of the original uranium nucleus. A complete fission
of one kilogram of uranium-235 produces a mass loss of a single gram. This lost
mass is then completely converted into energy. An energy mass of 24 million kilo-
watt hours is thereby released. Around 3000 tons of coal would have to be burnt to
release the same amount of energy.
After Hahn’s discovery the use of nuclear energy was promoted mainly by the mili-
tary. Albert Einstein, who emigrated to the USA in 1933 to escape Nazi persecution,
sent a letter to US president Roosevelt on 2 August 1939 warning him that Hitler’s
Germany was making a serious effort to produce pure uranium-235 that could be
used to build an atomic bomb. When the Second World War broke out on 1
September 1939, the American government set up the Manhattan Project with the
aim of developing and building an effective atomic bomb.

Germany as an Example of Nuclear History


The Paris Treaty of 5 May 1955 allowed Germany non-military use of
nuclear energy. Expectations for the nuclear industry ran high. A separate
ministry for nuclear energy was created, and the first minister was Franz Josef Strauss.
On 31 October 1957, Germany put its first research reactor, called the nuclear egg, into
operation at the Technical University in Munich. In June 1961 the Kahl nuclear power
station fed electricity into the public grid for the first time. In 1972 the Stade and
Wuergassen commercial nuclear power stations began to provide electricity, and with
Biblis the world’s first 1200 megawatt block went into operation in 1974. In 1989 the
last new power station, Neckarwestheim, was connected to the grid. Until that point the
federal government had invested over 19 billion euros in the research and development
of nuclear energy. However, public concerns about the risks of nuclear energy continued
to grow and prevented the building of new power stations. In 2000 Germany finally
renounced the use of nuclear power, following the example of Austria, Italy, Sweden
and Belgium. The last nuclear power station in Germany is scheduled to be disconnected
from the grid in 2023. Despite more than 50 years of nuclear energy use in Germany,
the problem of end storage for highly radioactive waste has still not completely been
resolved.

9
1 Our Hunger for Energy

The biggest problem turned out to be the ability to produce significant quantities of
uranium-235 to maintain a chain reaction. If metallic uranium is refined from
uranium ore, there is a 99.3% probability that it will consist of heavy uranium-235.
This is practically useless for producing a bomb. It even has the characteristic of
decelerating and absorbing neutrons, thus bringing any kind of chain reaction to a
halt. Only 0.7% of available uranium consists of uranium-235, which must be
enriched proportionally higher to create a chain reaction. No separation between
uranium-235 and uranium-238 can be achieved by chemical means because chemi-
cally both isotopes are totally identical. Consequently, other solutions had to be
sought. Ultimately, this separation succeeded through the use of a centrifuge because
the isotopes have different masses.
The Manhattan Project cost more than US$2 billion between 1939 and 1945. The
desired results were finally achieved under the direction of the physicist J. Robert
Oppenheimer: on 16 July 1945, two months after the capitulation of Germany, the
first test of the atomic bomb was carried out in the US state of New Mexico. Using
the bomb on Germany was no longer up for discussion, but shortly before the end
of the Second World War the atomic bomb was dropped on the Japanese cities
Hiroshima and Nagasaki – with the well-known aftermath.
The non-military use of nuclear energy came some years later. Although physicists
like Werner Heisenberg and Enrico Fermi had been conducting tests in reactors
since 1941, it was not until December 1951 in Idaho that the research reactor EBR
1 succeeded in generating electric current using nuclear energy.

■ www.iaea.org/programmes/a2/ IEA Power Reactor Information System


■ www.facts-on-nuclear-energy.info Nuclear Power Fact File Poster
Campaign

Unlike the uncontrollable chain reaction that occurs when an atomic bomb explodes,
nuclear fission in a nuclear power station should be controllable. Once a chain reac-
tion has started, the number of new neutrons resulting from the nuclear fission must
be kept to a limit. Each splitting of a uranium nucleus releases two to three neutrons,
only one of which is allowed to split another nucleus. Control rods that capture the
neutrons reduce the number of neutrons released. If this number is too high, the
process gets out of control. The nuclear power station then starts to act like an atomic
bomb and an uncontrolled chain reaction occurs. Technically, and this was the
leading view at the time, nuclear fission can be controlled and undesired reactions
eliminated.
The early euphoria that came with the use of nuclear energy died down when an
accident occurred with a reactor on 28 March 1979 in Harrisburg, the capital of the
US state of Pennsylvania. Large amounts of radioactivity escaped. Many animals
and plants were affected and the number of stillbirths among the nearby population
increased dramatically after the tragedy.
On 26 April 1986 another serious accident occurred with a nuclear reactor in
Chernobyl, a city in the Ukraine. What was thought to be impossible happened: the

10
1.1 Energy Supply – Yesterday and Today

chain reaction got out of control and the result was a nuclear meltdown. The radio-
activity that was released produced high radiation levels in places as far away
as Germany. Many helpers who tried to contain the damage on site paid for their
efforts with their lives and thousands of people died from cancer in the years that
followed.
Another problem with the civilian use of nuclear energy is the disposal of radioac-
tive waste. The use of uranium fuel elements in nuclear power stations produces
large quantities of radioactive waste that will create a deadly threat for centuries to
come. No safe way has yet been found to dispose of this waste.
Technically, the use of nuclear energy is fascinating and the prospect of generating
electricity from relatively small amounts of fuel is very tempting. But there are
serious risks involved. Germany has therefore agreed to a general decommissioning
of its nuclear energy plants. Once the last nuclear energy plant has been switched
off, the country’s adventure into this field will have cost the German federal govern-
ment alone more than 40 billion euros in research and development. Germany’s
most expensive leisure park has become a bizarre showpiece for the enormous bad
investment in nuclear energy. The prototype for a fast breeder reactor was erected
for around 4 billion euros in the North Rhine-Westphalian town of Kalkar. Due to
concerns about safety, including those relating to the highly reactive cooling agent
sodium, the nuclear plant was never put into operation. Today the Kernwasser
Wunderland Kalkar leisure park is located in the industrial ruins of the nuclear plant
(Figure 1.4).

Figure 1.4 The Kernwasser Wunderland leisure park is in the grounds of a fast breeder reactor in
Kalkar that was never put into operation. Photos: www.wunderlandkalkar.eu.

Whereas Germany reluctantly turned its back on nuclear energy after many years
of grappling with the potential it offers, other countries have taken a totally different
stance. One example is France, which uses nuclear power to supply about 80% of
its electricity needs. China is also enthusiastic about the use of nuclear and, due to

11
1 Our Hunger for Energy

the efforts of former US president George W. Bush, plans to continue the develop-
ment of nuclear energy are again on the table in the USA.
In 2007 there were 444 nuclear power stations in operation worldwide. Yet nuclear
energy is relatively unimportant for the global supply of energy. Its share is similar
to that of hydropower and much lower than that of firewood. If a major effort were
made to replace the majority of fossil power plants with nuclear energy, uranium
supplies would be depleted within a few years. In this sense, nuclear power stations
are not a real alternative when it comes to protecting the environment – although
this is how some politicians, and specifically the companies that would profit from
the use of nuclear power, often like to present the option to the public.
In the long term there are high hopes for a totally new variant of atomic energy:
nuclear fusion. The model for this technology is the sun, which releases its energy
through a nuclear fusion of hydrogen nuclei. The aim is to duplicate this process
on earth without the danger of triggering an undesirable chain reaction like
Chernobyl. But there is a hitch to this plan: the particles must be heated to tempera-
tures of several million degrees centigrade to initiate the momentum of nuclear
fusion. There is no known material that can permanently withstand such tempera-
tures. Therefore, other technologies, such as the use of strong magnetic fields to
contain reaction materials, are being tested. These technologies have seen some
success, but despite the enormous amounts of energy used for ignition, the reactors
always go out by themselves.
Currently there is no one who can seriously predict whether this technology will
ever actually work in practice. Critics point out that the proponents of nuclear fusion
have been saying for years that it will take 50 years for a commercial functioning
reactor to be connected to the grid. Despite the passage of time that 50-year time
frame never reduces – the only thing about nuclear fusion that can be said with any
certainty.
However, even if this technology became advanced enough to use, there are two
good reasons for opposing the development of nuclear fusion. Firstly, this technol-
ogy is decidedly more expensive than nuclear fission today. For economic reasons
preference would be given to alternatives such as renewable energies. Greater
investment in fusion testing means less investment in alternative energies. Secondly,
the nuclear fusion plants produce radioactive substances and waste that present
a risk.

1.1.5 The Century of Fossil Energy


Whereas traditional renewable energies covered most of the energy needs of
mankind until the end of the 19th century, the 20th century can be seen as the century
of fossil energy. By the middle of the century fossil fuels in internal combustion
engines had almost completely replaced the classic systems for using renewable
energies, such as windmills, water wheels and vehicles and machines driven by
muscle power. Modern hydropower for the generation of electricity and biomass,
which was mainly used as fuel, were the only renewable energies of any
significance.

12
1.2 Energy Needs – Who Needs What, Where and How Much?

After the Second World War the demand for energy soared, and fossil energy
sources were able to increase their share substantially. In 2007 fossil energies
covered around 80% of the world’s primary energy needs (Figure 1.5). Hydropower
and nuclear energy had a share of around 6% and 5%, respectively, and biomass
close to 10%. The other renewable energies amounted to less than 1% with geo-
thermal energy – use of the earth’s heat – having the largest share. At the beginning
of the 21st century renewable energies such as wind power and solar energy, which
until then had been used relatively little, started to record double-digit growth rates.
It is expected that their share of worldwide energy supply will increase considerably
in coming years.

550
other renewables
500 biom ass
hydro power
450
nuclear power
primary energy demand in EJ
_

400 coal
natural gas
350
crude oil
300

250

200

150

100

50

0
1965

1967

1969

1971

1973

1975

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

2005

2007
Figure 1.5 Development of primary energy demand worldwide since 1965.

1.2 Energy Needs – Who Needs What, Where and


How Much?

Demand for energy is distributed unevenly across the world. Six countries, namely
the USA, China, Russia, India, Japan and Germany, use more than half the available
energy.
The USA alone needs one-fifth of the energy used in the world, even though only
one-twentieth of the worldwide population lives there. If every citizen of India were
to use as much energy as each American, global demand for energy would rise by
about 70%. If all the people on earth developed the same hunger for energy as the
Americans, demand would increase fourfold.

13
1 Our Hunger for Energy

Energy Cannot be Used Up


Anyone who has taken physics at high school will have learned about the
concept of energy conservation. According to this principle, energy cannot
be consumed or produced but only converted from one form into another.
The car is a good example. The fact that cars consume too much is something we are
keenly aware of each time we fill up with petrol. The petrol that a car needs, and we
wince every time we pay for it, is a type of stored chemical energy. Combustion produces
thermal energy. This is converted by an engine into kinetic energy and transferred to the
car. Once all the petrol has been consumed, the car stops running. However, this does
not mean that the energy has disappeared. Instead, due to the waste heat from the engine
and the friction of the tyres and with the air, the energy has been dispersed into the
environment. However, this ambient heat cannot be used any more; we are unable to
produce petrol from ambient heat. When a car is driven, the usable energy content of
the petrol is changed into an ambient heat that is no longer usable. This means that this
energy is lost to us and thus consumed, even if this is not correct in terms of the laws
of physics.
A photovoltaic system is a different matter. It converts sunlight directly into electric
energy. It’s often said that a solar system produces energy. From the point of view of
physics, this is also incorrect. A solar system merely converts hard-to-use solar radiation
into high-quality electricity.

When discussing which countries consume the highest amounts of energy, it is


important to look beyond overall consumption figures. Population numbers also play
an important role in any comparison. In absolute terms, India consumes more energy
than Germany or the UK. But this is to be expected with a population of more than
one billion people. Consumption per head in India is less than one-seventh of that
in Germany or the UK. Although India is the country with the fourth highest use
of primary energy in the world, its consumption per head is less than half the world
average.

Primary Energy, Apple Energy and Orange Energy


If we compare our own electricity and gas consumption, we see that our
consumption will almost always be higher if we heat our homes with gas. A
comparison of the gas and electricity bills will not show much of a difference, though.
Electricity and natural gas are two types of energy or energy sources that, like apples
and oranges, cannot be compared directly like-for-like. Two to three kilowatt hours of
gas have to be burnt in a power plant in order to produce one kilowatt hour of electricity
from gas. The rest usually disperses unused into the environment as heat. When compar-
ing different forms of energy, a distinction is therefore made between primary energy,
secondary energy and useful energy.
Primary energy is energy in its natural and technically unconverted form, such as coal,
crude oil, natural gas, uranium, sunlight, wind, wood and cow dung (biomass).
Secondary energy is energy in the form in which it is channelled to users. This
includes natural gas, petrol, heating oil, electricity and district heating (the use of a
centralized boiler installation to provide heat for several buildings).

14
1.2 Energy Needs – Who Needs What, Where and How Much?

Useful energy is energy in its eventual form, such as light for illumination, warmth
for heating and power for machines and vehicles.
The different forms of energy are most frequently compared on a primary energy basis.
More than 90% of the original energy content is lost during the conversion of primary
energy to usable energy.

Figure 1.6 shows global primary energy needs per head. It is evident that the
Western industrialized states and countries with large supplies of crude oil have a
high rate of consumption. Prosperity and cheap energy prices boost consumption.
When it comes to the geographical pattern of consumption, the map clearly shows
that the countries with very high consumption – with the exception of Australia,
New Zealand and South Africa – are all in the Northern hemisphere. Germany and
France alone consume more than the entire African continent.

<25 % 150 % ... 200 %


25 % ... 50 % 200 % ... 300 %
50 % ... 150 % 300 % ... 400 %
>400 %

Figure 1.6 Primary energy usage per head related to the world average.

Countries with especially high energy consumption mostly use fossil energy sources
to satisfy their energy needs. On the other hand, countries with particularly low
energy needs rely to a large degree on traditional biomass. This includes firewood
and other conventional animal or plant products, such as dried animal dung. More
than two billion people worldwide use firewood and charcoal for cooking and
heating. In Africa south of the Sahara about 90% of the population is totally depend-
ent on fuels from traditional biomass.

15
1 Our Hunger for Energy

Big differences also exist between the industrialized countries, however. Whereas
many of them, such as Germany and the USA, use fossil fuels or nuclear energy to
cover more than 80% of their primary energy requirements, certain other industrial-
ized countries have already increased their share of renewable energy use consider-
ably. The countries of the Alps, Norway and Sweden use a noticeably high proportion
of hydropower. Biomass also plays a big role in some countries like Sweden and
Finland. In Iceland the natural heat of the earth is the energy form with the highest
share. Hydropower and geothermal energy together cover around 70% of Iceland’s
energy requirements.
Ethiopia, on the other hand, is a typical example of one of the poorest countries in
the world. More than 90% of the energy it uses is still based on traditional biomass.
Figure 1.7 shows the difference in how four countries use key forms of energy to
cover their energy needs.

natural gas
23 % nuclear
power
8% biom ass
hard coal biom ass 93 %
23 % 2.8 %
hydropower
2.7 % crude oil
6%
hydropower
crude oil 1%
40 % USA Ethiopia

geothermal Iceland Germany natural gas


energy 23 %
55 %
lignite nuclear
11 % 12.5 % biomass
3.3 %
wind
hydropower hard coal 0.7 %
16 % 13 %
hydropower
crude oil 0.5 %
26 % crude oil solar and
36 % geothermal
0.1 %
hard coal
3%

Figure 1.7 Percentage of different energy sources covering primary energy requirements in Ethiopia,
Germany, Iceland and the USA in 2005.

1.3 ‘Anyway’ Energy

According to statistics, only less than one percent of Germany’s primary energy con-
sumption was covered by solar energy in 2005. In the UK and the USA the percent-
age is even lower. The proportion of other renewable energies is also still quite low.

16
1.3 ‘Anyway’ Energy

This makes it difficult for most of us to imagine that renewable energies will be
riding to the rescue of the environment in a few years. In reality, however, renew-
able energies already constitute over 99% of German energy resources if one looks
at the complete picture of energy use.
Winston Churchill supposedly said ‘Do not trust any statistic that you did not fake’.
It is widely believed that fossil fuel sources cover the lion’s share of our energy
needs. At least this is what all the usual statistics on energy claim. But it is only
true if we define our energy requirements in a very narrow way.
The heat of a radiator, the light provided by a conventional light bulb and the driving
energy of a ship’s diesel engine generally form an integral part of our energy
requirements. What is not included in any statistics on energy is the warming effect
of the sunshine streaming through windows, the sunlight that illuminates houses
and streets so that artificial lighting can be switched off during daylight, and the
wind that can propel sailing boats right across the Atlantic. A heated greenhouse
that uses artificial light to grow useful plants is included in the statistics on energy;
on the other hand, a covered early planting of vegetables that uses only natural
sunlight is not included. The floodlight illumination of a stadium during an evening
football game falls under our energy needs. If the football game takes place in the
bright sunlight, the statistics on energy will claim that the football arena that is
brightly lit up by the sun actually does not need any light. If we switch on snow
blowers to compensate for the ever-decreasing amount of snow available in ski
areas, this becomes a case for the statistics, whereas natural snow is not. When we
fill our drinking water storage containers using electric pumps, we have to pay for
the energy used. If rain fills the storage containers, this is not considered in the
statistics. The high amount of electricity needed to run electric dryers also increases
energy use. On the other hand, if the washing is dried by the wind and the sun on
a conventional clothesline, this does not constitute an energy need as far as the
statistics are concerned.
Natural and technically unconverted forms of energy are not a component in our
energy requirements in a conventional sense. Yet it should not make any difference
where we derive the energy needed to heat our bath water, grow our plants or
provide light. We take the availability of natural renewable energy forms such as
solar energy so much for granted simply because they are there anyway and thus
appear to have so little value that they do not even merit mention in the statistics.
However, this distorts our impression of our energy requirements and puts the pos-
sibilities of renewable energy in a false light. This can be illustrated using the
example of energy consumption in Germany.
Germany covers an area of 357 093 km2 and the annual solar radiation is on average
1064 kilowatt hours per square metre. Germany therefore benefits from 380 trillion
kilowatt hours of energy from the sun each year. This is about 100 times as much
as the primary energy consumption recorded in the statistics for Germany and even
more than the entire primary energy needs of the world. Part of this radiation heats
the earth and the air; another part is converted into plant growth, thus producing
biomass.

17
1 Our Hunger for Energy

Around 800 millimetres or 0.8 cubic metres of rain fall per square metre in Germany.
The annual rainfall for all of Germany adds up to 286 billion cubic metres. The sun
evaporates this water before it reaches earth in the form of rain. One cubic metre
of water requires 627 kilowatt hours to evaporate. This means the annual rainfall
contains around 170 trillion kilowatt hours of energy.
About 2% of solar energy is converted through the movement of the wind. In
Germany this amounts to around 8 trillion kilowatt hours. Sun, wind and water
together produce abound 567 trillion kilowatt hours of energy in Germany each
year. Geothermal and ocean energy are not even included in these figures. If this
quantity of energy were to drop even a few percentage points, the result would be
drought and arctic winters (Figure 1.8).

anyway anyway wind


hydro 8 PWh
179 PWh

nuclear
0.5 PWh
fossil fuels
3.3 PWh

anyway sun
380 PWh Figure 1.8 Total energy resources in Germany
taking into account ‘anyway’ energy; that is,
natural renewable forms of energy.

The statistics of 2007 listed Germany’s primary energy requirements as being


around 14 000 petajoules. This converts to close to 4 trillion kilowatt hours. Of
course, the statistic includes solar energy, hydropower and wind energy. The propor-
tion of all renewable energies combined in the requirement for primary energy totals
0.183 trillion kilowatt hours. This is the amount that technical installations convert
into renewable energy. The natural forms of renewable energy that exist anyway
are totally omitted from this statistic. This explains the small obvious statistical
discrepancy for the previous calculation of 567 trillion kilowatt hours for renewable
energy resources. The difference between this and conventional statistics is clarified
in Figure 1.8, where the natural renewable forms of energy that previously were not
recorded in the statistics are referred to as ‘anyway’ energy – in other words, energy
that exists anyway.
Anyone who thinks that these calculations amount to statistical hairsplitting is
wrong. As the facts about climate change have now become public knowledge, there
is a general interest in replacing fossil fuels with renewable energies as quickly as
possible. But many people are under the impression that this is difficult to implement
and almost impossible to accomplish within a reasonable period of time. The claim
that solar energy only constitutes an insignificant share of energy resources is
repeated like a prayer wheel. If the claim were true, this scepticism would be justi-
fied. In fact, it is fossil and nuclear energies that make up 0.6% of the energy
resources in Germany. And no one could seriously doubt that 0.6% is replaceable
in the foreseeable future.

18
1.4 Energy Supplies – Wealth Forever

The eruption of the Tambora volcano in Indonesia in 1815 shows us what happens
when even a fraction of ‘anyway energy’ is lost. The gigantic quantities of volcanic
gases and dust that were emitted into the atmosphere reduced the amount of sunlight
available in the following years. In 1816 and 1817 Europe experienced massive crop
failures. Ten thousand people died of starvation. If something like this happened
today, we would suffer similar consequences. A large proportion of the energy that
safeguards our food supply comes from a natural source – the sun. The small and
diminishing supply of fossil and nuclear energy could not come anywhere close to
compensating for even relatively minor fluctuations in natural energy.
The question then is what is ‘anyway energy’ worth? In Europe in 2007 oil cost
more than 4 cents (euro) per kilowatt hour before tax; natural gas was 2 cents and
rising. Because solar radiation and wind power cannot be stored as easily as oil and
natural gas, their value should be set at half that of natural gas, thus 1 cent per kilo-
watt hour. Hydropower, on the other hand, would be set at 1.5 cents per kilowatt
hour because it is easier to store. The total value of ‘anyway energy’ can thus be
valued at about 6.5 trillion euros per year. According to this calculation, anyway
solar energy alone is worth around 4 trillion euros. And in the USA it is worth the
equivalent of 100 trillion euros.
Natural renewable forms of energy in the order of 567 trillion kilowatt hours are
not recorded as a separate entry in statistical calculations in Germany. This means
that the public perception of the energy supply is distorted. We are left with the
false impression that fossil and nuclear energy sources make up the major portion
of our energy supply. In reality, the share from these sources is less than 1% and
we should be replacing this with renewable energies as soon as possible to protect
the climate. Natural renewable forms of energy to the value of around 6.5 trillion
euros are available to us today free of charge. We cannot afford to ignore this
option.

1.4 Energy Supplies – Wealth Forever

When we use fossil energy sources today, we are utilizing solar energy that was
stored millions of years ago – but without the possibility of renewing this source in
the foreseeable future. Yet our current hunger for energy is so great that most of
the known fossil deposits will be used up during the 21st century. And suitable
deposits of fuel uranium for conventional nuclear plants are also becoming rarer.
For decades pessimists have been warning about the imminent end to fossil energy
reserves. Yet this end never quite seems to be in sight, and most people take no
heed of the warnings. It was not until oil prices started their steep spiral in 2000
that the message that ‘black gold’ would one day run out sank home.
In the past constant technological advances in the exploitation of oil and natural gas
have always resulted in a revision of the forecasts about how long reserves would
last. The large coal reserves still available worldwide, in particular, could enable us
to use fossil energy sources for decades or even another century.

19
1 Our Hunger for Energy

But the number of new finds, specifically of oil, has declined substantially during
recent years, and new supplies cannot be exploited fast enough to meet rising
demand. In the long term oil prices will therefore continue to rise, even if brief dips
in prices seem to signal an easing of the situation. On one hand demand is rising,
whereas supplies tend to be dwindling; and on the other hand the effort needed to
exploit new supplies is increasing along with the costs.
During the first commercial drilling in America in 1859 oil could be found at depths
of 20 m, whereas today drilling depths of up to 10 000 m are no longer uncommon.
Significant technical progress has also been made in locating possible deposits, and
so we know far more today about possible finds than we did several decades ago.
However, this also makes it highly unlikely that any major new finds will be
discovered.
At the current rate of production the known supplies in the USA and Great Britain
will be depleted in about a decade. This increases the dependency, particularly of
the industrialized nations, on a small number of producing countries. More than
60% of extractable oil supplies are found in the Middle East (Figure 1.9). The
biggest oil producers in the region are Iraq, Iran, Kuwait, Saudi Arabia and the
United Arab Emirates.

15
3

14 6

100 Gt
14

assumed
already reserves still additional
82 Gt
exploited available (in Gt) global
resources

Figure 1.9 Distribution of oil reserves on earth by region. Status 2004.

This region has been the scene of major conflicts in recent years, and its large oil
reserves are likely to increase tensions even further in the future. The dependency
of the industrialized nations on the OPEC countries will also increase because these
countries have almost three-quarters of the known reserves.
The current extent of availability can be calculated by dividing the known exploit-
able reserves by current production. In the case of oil, this is about 43 years (see

20
1.5 The End of Fission

Figure 1.10). This availability could drop further if there is an increase in annual
production. New deposits are being exploited in addition to known supplies. It is
estimated that the reserves will increase up to one-half times due to these additional
supplies. If production remains constant over the next few decades, the oil reserves
will last another 65 years. However, it is certain that we will no longer be using oil
as an energy source at the end of this century.

hard coal 207

lignite 198

natural gas 64

crude oil 43

uranium 42

0 50 100 150 200 years 250

Figure 1.10 Extent of availability (in years) of known energy reserves based on current
production.

The situation with natural gas and coal supplies is not quite so critical. Based on
current production levels, the known gas supplies will be depleted in 64 years. In
contrast to oil, the estimated additional supplies are considerably more extensive
than those known to date. This is partly because the deposits are located at a lower
depth than oil and also because industrial production and the search for new supplies
began much later. However, due to the continuing high level of consumption, natural
gas supplies will also be running out during this century. Coal is the only fossil fuel
that may still be available at the dawn of the next century.

1.5 The End of Fission

A key point about fuel supply, and one that most people are unaware of, is that even
uranium supplies are very limited. Although there is more uranium in the earth’s
crust than either gold or silver, less than one percent of the purest natural uranium
can be used to create energy. Power stations can only use natural uranium after the
useable part of uranium has been enriched with uranium-235.
The share of ore must be higher than average to enable an effective exploitation of
natural uranium. Canada is the only country that has deposits with a uranium ore
content of more than one percent. If the uranium ore content drops, considerably

21
1 Our Hunger for Energy

larger amounts of it have to be mined for its degradation. This substantially increases
the energy required to extract the ore and contributes to the costs.
Despite intensive efforts by some countries to develop nuclear energy, at 6% its
share of worldwide primary energy supply is still relatively low. A few countries
like France are using nuclear power stations to cover 80% of their electricity needs.
However, even in France cars cannot be run on nuclear power, and only some houses
use nuclear energy for heating. Consequently, nuclear energy only really constitutes
40% of the total primary energy supply, even in France.
The supplies of uranium would be depleted in just a few years if nuclear energy
were used to replace all fossil energies, which would also mean developing nuclear
automobiles and nuclear heating. Power stations could use other techniques, such
as relatively risk-free fast breeders, to increase the amount of energy they are able
to exploit, but this would do very little to change the fact that the supply of uranium
is limited. Even the uranium supplies that can be exploited economically will run
out in a few decades at the most – which does not make a convincing argument for
building new nuclear power stations with a lifespan of 30 to 40 years. For these
reasons alone nuclear energy is not a viable alternative to fossil energies.

1.6 Oil Prices Today – Politics, Supply and Demand

The low energy prices of the 1970s were the foundation not only of the economic
miracle in several industrial countries, but of the massive rise in energy consump-
tion. The founding of OPEC and the politically motivated limiting of production in
1973 led ultimately to a dramatic increase in oil prices. The shocked industrialized
countries reacted with relative helplessness. In 1973 they founded the International
Energy Agency (IEA) to coordinate their energy policies and ensure that the supply
of energy remained secure and affordable.
The purpose of strategic oil reserves is to guarantee availability when the oil supply
is interrupted and to stabilize prices. For example, Germany stockpiles 25 million
tons of crude oil or crude oil products that can cover the country’s oil requirements
for 90 days. The US strategic petroleum reserve is the largest in the world and holds
up to 99 million tons.
The commitment to developing the use of renewable energies also increased in the
1970s. However, a large number of failed mammoth projects showed that cost-
effective and sustainable energy supply is not something that can be forced through;
it can only be the outcome of long-term, ongoing development. Nevertheless, the
oil crisis in the 1970s paved the way for the current boom in renewable energies.
The 1990s were marked by extremely low oil prices. As a result, efforts to save
energy and to develop renewable energies stagnated. Due to booming global eco-
nomic activity and extremely high demand, especially from China, oil prices reached
new heights after the year 2000. Oil prices in 2006 were almost double what they
were at the time of the oil crises in the 1970s (see Figure 1.11). Up to now this has
only had a limited effect on the world economy. This can be explained by consider-

22
1.6 Oil Prices Today – Politics, Supply and Demand

100

__
in $/barrel, $2000/barrel and €2000/barrel
$/barrel, nominal
90 €/barrel, adjusted f or inf lation
80 $2000/barrel $/barrel, adjusted f or inf lation

crude oil prices 70

60
€ 2000/barrel
50

40

30

20
$/barrel
10

0
1972
1974
1976
1978
1980
1982
1984
1986
1988
1990
1992
1994
1996
1998
2000
2002
2004
2006
2008
Figure 1.11 Development of oil prices in current prices and inflation-adjusted.

ing the inflation-adjusted oil prices. In 1980 one US dollar bought twice as much
as it could buy in 2005. To this extent, the inflation-adjusted oil price at the time
was also twice as high. Another reason is that the economy today depends consider-
ably less on energy prices than at the time of the oil crisis. However, an unreasonably
high oil price would have a massive impact on the world economy.
As the supplies of fossil energies begin to run out, oil, natural gas and coal prices
will rise further. The dip in prices resulting from the financial crisis in 2008 at best
offers a short breather. Another round of price increases is a certainty. Political risks
and a growing reliance on certain countries rich in raw materials also conceal the
considerable possibility of another sudden increase in prices. For economic reasons
it is important that some urgency be given to developing an alternative energy
supply beyond fossil or nuclear energy sources.
During the period of transition supply and demand will also allow a fluctuation in
the prices of renewable energies, as was shown by the price increase for wood-
burning fuels in 2006. However, in the long term the prices for renewable energies
will continue to drop as a result of ongoing technical advances and more efficient
production, whereas the prices for fossil energy sources and nuclear energy will
continue to rise.

23
2
2 The Climate Before
the Collapse?
We have known for a long time that the climate is changing. Numerous ice ages
and warming periods have shown that the climate on earth is constantly undergoing
change. In terms of human lifetimes, each of these periods lasted a relatively long
period of time. In the more recent history of the earth, ice ages have occurred about
every 100 000 years, always interrupted by much shorter warming times. Our current
warming period, called the Holocene era, began about 11 700 years ago. As the
previous warming periods on average lasted only about 15 000 years, we should be
heading towards the next ice age.
The exact causes for the alternation between warming and ice ages can only be
reconstructed to a point. Natural effects, such as changes in solar activity, changes
in the geometry of the earth’s orbit, vulcanism, changes in the ocean currents and
the shifting of continental plates are considered the main causes of climate change.
When multiple causes occur at the same time, very abrupt changes are possible.
This explains the climate changes of the earth. In this sense, the global warming
that we have observed in recent years is nothing out of the ordinary. What is unusual
is that apparently for the first time it is living things on earth – namely, human
beings – that are causing an abrupt change in climate.

2.1 It Is Getting Warm – Climate Changes Today

2.1.1 The Ice is Slowly Melting


Climate conditions on earth have been relatively constant for several thousand years.
It is on this basis that our civilization, the places where we have settled and our
agricultural regions have established themselves. During the last ice age areas that

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

24
2.1 It Is Getting Warm – Climate Changes Today

today are densely populated would have been covered with thick layers of ice. And
during the hottest periods in our climate’s history our coastal cities would have been
sunk metres deep below the ocean. If major changes to climatic conditions occur,
they will undoubtedly have an even stronger impact on the face of the earth and our
current conditions of life than any of the most dramatic historical events of the last
few centuries.

Observed Climate Changes (Intergovernmental Panel on


Climate Change, 2007)
■ The global surface temperature rose by 0.74 °C between 1906 and 2005.
■ Eleven of the last 12 years have been the warmest since records were kept.
■ The increase in temperature during the last 50 years was twice as high as during the
last 100 years. The warming of the Arctic has been twice as fast.
■ The temperatures of the last 50 years have probably been higher than at any time in
the last 1300 years.
■ Glaciers are shrinking worldwide, as are the ice sheets on Greenland and
Antarctica.
■ Since 1993 the sea level has risen an average of 3.1 mm per year; during the 20th
century this amounted to a total of 17 cm. More than half of this is due to the thermal
expansion of the oceans, about 25% to the melting of mountain glaciers and around
15% to the melting of the arctic ice sheets.
■ The frequency of heavy precipitation has increased.
■ The frequency and intensity of droughts has increased since the 1970s.
■ The frequency of extreme temperatures has increased.
■ Tropical cyclones have become much more intense since the 1970s.

During the past 100 years the average temperature on earth has increased by a good
0.7 °C. At first glance this does not seem like much. However, this warming is not
occurring at a uniform rate in all regions and is also not constant throughout the
year. If temperatures fall into a normal range for ten months, then two months can
follow at more than 4 °C above the average. In some regions of the world the tem-
perature is already higher by 2 °C as an annual average (Figure 2.1).
As a result of global warming, the water in the oceans expands. Due to the rise in
temperatures, more and more Arctic ice and the permanent ice of the glaciers also
melt. As a result, the sea levels rise even higher.
The temperatures in the polar region rise even more quickly than in the rest of the
world. The ice coverage in the Arctic has decreased by about 10 to 15% within 20
years (Figure 2.2). Along with the ice masses of the Arctic, many glaciers are also
melting at a fast rate. The largest glacier in the world, the Bering Glacier in the
Arctic region of Canada, has shrunk by more than 10 km during the past century.

25
2 The Climate Before the Collapse?

Figure 2.1 Temperature change from 2002 to 2006 compared to the long-standing average.
Source: NASA, svs.gsfc.nasa.gov.

Of the mountain glaciers in the Eastern Alps less than half the mass that existed in
1850 remains today.

2.1.2 Natural Catastrophes Occur More Frequently


As global temperatures rise, extremes in weather are also occurring. Major differ-
ences in temperatures cause storms to be more violent, rainfall to be heavier and
high tides and flooding to be more frequent than before. The Munich Reinsurance
company in Germany has long been concerned about the increase in natural catas-
trophes and resulting damage. In the event of loss or damage, reinsurance companies
are responsible for covering part of the insured damage.
Economic as well as insured damage has risen sharply since the 1950s, as shown
in Figure 2.3. Yet the graphics only include major catastrophes. Small local and
medium-sized events are not included in the chart. There are two reasons for the
rise: firstly, due to growing prosperity in the world, there is an increase in the number
of valuable things that can be damaged when a natural catastrophe occurs. Secondly,
the frequency and intensity of natural catastrophes has increased considerably.
Hurricane Katrina alone, which devastated the city of New Orleans in 2005, caused
around US$125 billion worth of damage and cost the lives of 1300 people (also see
Figure 2.4).
Europe has also experienced an increase in extreme and devastating weather condi-
tions. Examples include the 2005/2006 winter, which was the mildest since records
began, and the record heat in the summer of 2003. Major heatwaves such as this
one cause a decline in harvest yields.

26
2.1 It Is Getting Warm – Climate Changes Today

Figure 2.2 Arctic ice coverage in September, calculated for the years 1979 to 1981
(above) and for the years 2003 to 2005 (below). Source: NASA, svs.gsfc.nasa.gov.

160
bill.€ total damage
140 insured damage
(in 2006 values)
120 trend of total damage

100

80

60

40

20

0
1950 1955 1960 1965 1970 1975 1980 1985 1990 1995 2000 2005

Figure 2.3 Total damage and insured damage due to major natural catastrophes
worldwide. Data: Munich Re Group (Quaschning, 2007).

27
2 The Climate Before the Collapse?

Figure 2.4 Damage caused by hurricanes in the USA. Photos: Munich RE Group.

Examples of Major Natural Catastrophes


■ Winter 1990: The storms Daria, Herta, Vivian and Wiebke kill 272
people in Europe and cause 12.8 billion euros in damage.
■ 29 April 1991: A storm tide resulting from the tropical cyclone Gorky hits
Bangladesh. 138 000 people die. The material damage at 3 billion euros is compara-
tively low in this poor country.
■ 12 December 1999: Storm Lothar devastates large areas of Europe. 110 people die.
The damage amounts to 11.5 billion euros.
■ August 2002: Unusually heavy rainfall with up to 400 litres per square metre causes
major flooding. In Germany, Dresden is one of the cities overcome by floods. In
Europe overall 230 people lose their lives and the damage is 18.5 billion euros.
■ August 2003: Europe’s most extreme heatwave in recorded history takes 70 000
lives and causes damage close to 13 billion euros.
■ August 2005: Hurricane Katrina wreaks havoc in the USA and destroys the city of
New Orleans. 1322 people die as a result. The most expensive storm of all time
causes US$125 billion (around 100 billion euros) worth of damage.
■ September 2005: Four weeks after Katrina, Hurricane Rita causes around 13 billion
euros worth of damage. About 3 million people are evacuated to protect the
population.
■ 18 January 2007: The storm Kyrill moves across Europe. The German railway halts
all train travel in Germany for the first time in history.

Extreme weather events also bring about a rise in the death rate. In the summer of
2003 about 70 000 more people died in Europe than in a normal year, as a result of
the excessive heat.
Although the financial damage resulting from natural catastrophes is still manage-
able at the moment in Europe, these figures are expected to rise dramatically by the
end of the century. The German Institute for Economic Research DIW projects that

28
2.2 The Guilty Parties – Causes of Climate Change

the overall cost to Germany alone for climate change will amount to around 3000
billion euros by the year 2100 (Kemfert, 2007) if global warming brings about a
4.5 °C rise in temperature.

2.2 The Guilty Parties – Causes of Climate Change

2.2.1 The Greenhouse Effect


Without the protective influence of the atmosphere, earth’s temperature would be
around −18 °C. This means we would be living on an ice-bound planet. Various
natural trace gases in the atmosphere – such as steam, carbon dioxide and ozone
– prevent the earth from emitting all incoming solar energy back into the universe.
As in a greenhouse, these gases radiate part of this energy back to earth. This natural
greenhouse effect is the basis for life on earth. As a result, the average temperature
has settled at around +15 °C. During the last few millennia a balance has been
created in the level of trace gases in the atmosphere, which has enabled life in the
form that we know it today.
Many different possible causes of observed climate change have already been dis-
cussed. For a long time sceptics even questioned whether climate change was really
taking place. Now that no one can really seriously claim that it is not getting warmer,
there are some who are trying to push the blame onto natural effects – for example,
on solar activity. Apparently solar activity during the last few decades has been
higher than in all the previous 8000 years. It has been demonstrated that the amount
of radiation that reaches the earth has indeed risen slightly. However, scientists rule
out the possibility that this is what is causing the high level of warming today. At
most, one-tenth of the observed increase in temperature is attributed to the rise in
solar activity (Figure 2.5).

Figure 2.5 Changes in solar activity


are responsible for only a fraction of
global warming. Picture: NASA.

29
2 The Climate Before the Collapse?

The most plausible cause of warming is that the proportions of the trace gases have
changed significantly due to human influences. The concentration of gases that have
been proven to cause global warming has increased considerably in recent decades.
Therefore, man is causing an increase in the natural greenhouse effect. The green-
house effect brought about by man is also referred to as an anthropogenic green-
house effect (Figure 2.6). But this is not a particularly new theory.

Figure 2.6 Causes of anthropogenic greenhouse effects due to human activities.

2.2.2 The Prime Suspect: Carbon Dioxide


In 1896 the Swedish scientist and Nobel prizewinner Svante Arrhenius already
worked out that doubling the carbon dioxide content (CO2) in the atmosphere
(Arrhenius, 1896) would increase the temperature by 4 to 6 °C. The connection
between observed climate warming and the increase in carbon dioxide following
industrialization was already being discussed in the 1930s. But at the time there was
no way of confirming this connection. It was not until the late 1950s that scientists
were able to prove that the concentration of carbon dioxide in the atmosphere was
actually increasing (Rahmstorf and Neu, 2004). Today there is widespread proof
that the increase in carbon dioxide concentration is one of the main causes of
observed warming.

Are We Ruining the Air that We Breathe?


The air that we breathe out contains around 4% carbon dioxide – about 100
times more than the air we breathe in. Every year, we each release around
350 kg of carbon dioxide into the atmosphere. When we light a campfire using wood,
we are also guilty of releasing carbon dioxide. However, plants, animals and people are

30
2.2 The Guilty Parties – Causes of Climate Change

all part of a biogeochemical cycle. People take in carbohydrates and breathe in oxygen.
They convert both substances into carbon dioxide, which they then breathe out again.
Plants in turn bond this carbon dioxide and provide our carbohydrates. Carbohydrates
are organic compounds consisting of carbon, hydrogen and oxygen and are produced in
plants through photosynthesis. For example, grains and noodles comprise 75% carbohy-
drates. It is even possible that the wheat in Italian spaghetti has converted the carbon
dioxide that we breathed out during our last Tuscan holiday into carbohydrates.
When a plant burns, rots or ends up as a carbohydrate supplier, it generates just as much
carbon dioxide as it previously extracted from the air. The natural cycles are therefore
CO2-neutral and do not lead to an increase in concentration. However, this does not
apply to the holiday trip to Italy and the transport of Italian spaghetti to other
countries.

Fossil energy use is the main cause of the increase in carbon dioxide concentration.
If we burn fossil energy sources, this is regarded as oxidation from a chemical point
of view. This reaction releases heat. We therefore use the effect of heat that results
when the carbon from oil, natural gas and coal is combined with the oxygen from
the air. The waste product we get in return is carbon dioxide – and we get it in
enormous quantities: currently over 25 billion tons. The average human produces
about 4000 kg per year. A corresponding amount of carbon dioxide would fill a cube
with sides 13 m long, or around two million one-litre bottles.
As with energy consumption, the emissions in individual countries can vary a great
deal (Table 2.1). For example, whereas someone from Mozambique would tip the
scales at less than 100 kg, thus one-tenth of a ton of CO2, per year, in China it is
almost 4 tons per head. In Germany it is 10 tons, and in the USA just short of 20
tons. If the carbon dioxide generated by Germans each year were dispersed over the
country’s entire land area, every German would sink down into more than one metre
of CO2. In contrast, the carbon dioxide in Mozambique dispersed over the country
would not even form a layer one millimetre thick.

Table 2.1 The ten countries on earth with the highest energy-related carbon dioxide emissions.
Status: 2006. Data: IEA (International Energy Agency, 2008).

Country Mil. Mil. t CO2/ Country Mil. Mil. t CO2/


t CO2 inhab. inhab. t CO2 inhab. inhab.
1. USA 5697 299 19.00 6. Germany 823 82 10.00
2. China 5606 1312 4.27 7. Canada 539 33 16.52
3. Russia 1587 143 11.14 8. Great Britain 536 61 8.86
4. India 1250 1110 1.13 9. South Korea 476 48 9.86
5. Japan 1213 128 9.49 10. Italy 448 59 7.61
World 28 003 6546 4.28 133. Mozambique 2 20 0.08

31
2 The Climate Before the Collapse?

Yet it is not that long since we have been able to claim with absolute certainty that
the amount of carbon dioxide in the atmosphere is increasing on a yearly basis. The
Mauna Loa Observatory in Hawaii has only been measuring the concentrations of
carbon dioxide continuously since 1958. At that time the concentration amounted
to 315.2 ppm, the following year it was already up to 315.8 ppm. The unit ppm stands
for ‘parts per million’. Thus there were 315 parts of carbon dioxide per million parts
of air. The small increase during the first year could also have been caused by a
measurement error or natural fluctuations. It was not clear until the values began
rising constantly in subsequent years that the amount of carbon dioxide was increas-
ing – and at a growing rate. In 2008 the CO2 concentration had already risen to
386 ppm.
But, in comparison with the enormous size of the atmosphere, the high carbon
dioxide emissions that occur when fossil fuels are burnt are minute. Furthermore,
some of the carbon dioxide is absorbed again by the oceans and the plants. So the
question is how much our emissions can even change the composition of the
atmosphere.
If we were to produce nitrogen instead of carbon dioxide when we use fossil energy
sources, this would not pose a big problem. The reason is that our air consists of
around 78% nitrogen and 21% oxygen but only 1% other gases – of which carbon
dioxide makes up only a minuscule part. During the course of the earth’s history,
the composition of the air has by no means been consistent. However, in the past
few thousand years a balance of less than 300 ppm has established itself. The pro-
portion of carbon dioxide in the atmosphere was thus less than 0.03%. However,
this is also the reason why we can even bring about any relevant changes. Small
quantities can be increased comparatively easily.
A study of the climate of the past few millennia requires a different approach. The
polar and Alpine ice sheets on earth have stored the history of the planet’s climate.
In the regions with permanent ice, fresh snow falls on the ice surfaces every year.
Large quantities of air are trapped between the ice crystals. The new snow masses
that form each year alongside the old ones increase the pressure on the old snow
and ultimately press it down into pure ice. In the process the air does not escape
totally but instead remains as small bubbles trapped in the ice. Today these air
bubbles can be examined using modern techniques of analysis. The deposits of snow
and the creation of ice repeat themselves each year with a regularity welcomed by
scientists. All it takes is to drill a hole into the ice and draw up ice samples from
deep down. This provides a silent witness to the past. The deeper one probes into
the ice, the further back in history one is able to look.
Different drill core studies are unanimous in their finding that the concentration of
carbon dioxide before the period of industrialization was around 280 ppm (Figure
2.7). But even the theory that high concentrations of carbon dioxide are a recurring
phenomenon could be refuted. The studies showed that the proportion of carbon
dioxide in the atmosphere is higher today than at any other point in the past 650 000
years (Intergovernmental Panel on Climate Change, 2007).
Once it was finally proven that carbon dioxide emissions are increasing, climate
models were developed to establish a correlation between the burning of fossil fuels

32
2.2 The Guilty Parties – Causes of Climate Change

ppm ppm
400 400
380 380
360 360
340 340
CO2 concentration

CO2 concentration
320 320
300 300
280 280
260 260
240 240
220 220
200 200
180 180
160 160
-400,000 -300,000 -200,000 -100,000 year 0 1000 1200 1400 1600 1800 2000

Figure 2.7 Development of carbon dioxide concentration in the atmosphere over the last 400 000
years and in the recent past. Data: CDIAC, cdiac.ornl.gov.

and the increase in CO2. Sources other than the emissions caused by humans were
not considered as reasons for the increase. The models show that the CO2 concentra-
tion could still more than double, depending on the extent of the future consumption
of coal, oil and natural gas.
This discovery raised another question, namely what the consequences of such
drastic changes would be. In high concentrations carbon dioxide is unhealthy for
humans and can even be life-threatening. However, to reach a level that would
damage the health of humans, the concentration would have to increase a hundred-
fold; to cause the danger of suffocation it would have to increase by a factor of 300.
There is no way that we could reach such levels, even if we were to burn all the
oil, natural gas and coal available. At least any acute danger to the lives of humans
can be ruled out.
Direct measurements of temperature have existed for well over 100 years. A com-
parison of global temperature changes with energy-related CO2 emissions shows a
significant connection between the two (Figure 2.8). Yet sceptics find fault with this
argument and point out that the temperatures actually fell slightly between the 1940s
and the 1980s. Today there is an explanation for this. Aerosols from high dust and
soot emissions in the burning of fossil energy sources reduce the solar radiation on
earth. They act, so to speak, like sunglasses and thus create a cooling effect. Today
modern filtering techniques have eliminated a major proportion of these emissions.
This again decreases the sunglasses effect. However, CO2 and the associated rise in
temperature have remained.
As wood absorbs carbon dioxide during its growth, its burning is carbon dioxide-
neutral. However, this is only the case if the same quantity of plants is used and
burnt as can grow back again. Tropical rainforests are currently being burnt at a
much faster rate than their rate of regeneration. An area of woodland the size of the
United Kingdom is disappearing every 18 months. Some of the forests being burnt

33
2 The Climate Before the Collapse?

30000 0.9
Mt °C

CO2 emissions from fuel combustion _ 25000 0.7

global temperature change


20000 0.5

temperature
15000 0.3

10000 0.1

5000 -0.1

CO2 emissions

0 -0.3
1860 1880 1900 1920 1940 1960 1980 2000

Figure 2.8 Progression of energy-related CO2 emissions and global changes in


temperatures since 1860.

are so large that they can easily be detected from space. If this does not change
during the next few years, almost all the woodland in the world will be gone in
about 100 years. During this time large quantities of carbon dioxide will be released,
causing 10% of the greenhouse effect. However, the remaining CO2 originates
largely from the burning of fossil energy sources.

2.2.3 Other Culprits


Fossil energy use is not the only culprit. Agriculture, the clearing of rainforests and
industry are also responsible for the additional greenhouse effect caused by people
(see Figure 2.6).
Aside from carbon dioxide, there are other substances released through the actions
of people that cause the greenhouse effect. Methane, hydrofluorocarbons (HFCs),
ozone and nitrous oxide are some of the key ones. Although the concentration of
these elements in the atmosphere is considerably lower than that of carbon dioxide,
they have a much higher specific greenhouse potential.
The greenhouse potential of methane is 23 times that of carbon dioxide. This means
that 1 kg of methane causes just as much damage as 23 kg of carbon dioxide.
Methane is the second most important greenhouse gas. Through the influence of
man the concentration of methane in the atmosphere has already more than doubled.
Methane is also emitted during the exploitation of fossil energies. It escapes
during natural gas extraction and coal mining, as well as from defective natural
gas lines. Refuse disposal sites are another source for the emission of methane
(Table 2.2).

34
2.2 The Guilty Parties – Causes of Climate Change

Table 2.2 Characteristics of the most important greenhouse gases. (Status 2008)

Carbon Methane HFC Ozone Laughing


dioxide close to gas (nitrous
earth oxide)
Chemical notation CO2 CH4 diverse O3 N2O
Concentration in the atmosphere in ppm 385 1.77 <0.01 0.034 0.32
Concentration in the year 1750 280 0.75 0 0.025 0.270
Annual increase in concentration + 0.4% +0.4% Varies +0.5% +0.25%
Greenhouse potential compared to CO2 1 23 >1000 >1000 296
Duration in the atmosphere in years 5-200 12 Varies <0.1 114
2
Reflection in W/m 1.66 0.5 0.35 0.35 0.16
Percentage of anthropogenic 56% 16% 11% 12% 5%
greenhouse effect

Interestingly, agriculture releases the largest proportion of methane. Livestock


rearing, decaying biomass and rice cultivation are the farming activities that produce
the most methane. Unlike carbon dioxide, it is not integrated into a biological cycle.
When ruminants digest green fodder, they create methane which they belch into the
atmosphere. A single cow or a herd would not pose a problem for the climate, but
the number of cattle worldwide is now so large that they represent a real threat to
the environment. Meat consumption per head has increased dramatically during the
past few decades, as has the number of people, so that there are now more than 1.3
billion cattle on earth.
Other greenhouse gases such as nitrous oxide (laughing gas) are created through
farming due to an excessive use of nitrogeneous fertilisers. Overall, agriculture
contributes around 15% of the greenhouse effect caused by man.
Other key greenhouse gases include ozone, fluoride hydrocarbons (HFC and halons),
steam and sulphur hexafluoride (SF6). Air pollutants help to create low-level ozone.
These pollutants originate in part from motorized traffic and, therefore, as a result
of the burning of fossil fuels. The ozone problem has become a familiar one because
of the health risk associated with summer smog. However, what is less known is
the impact of the ozone on the greenhouse effect. SF6 is used in the electricity sector
but the amount causing a greenhouse effect is relatively minimal.

Will the Hole in the Ozone Layer Keep Making Earth Warmer?
The ozone hole and the greenhouse effect are both global environmental
threats. Yet they have very little relationship to one another.
With the ozone a distinction must be made between the ozone layer at high altitudes and
ozone at a low level. The natural ozone layer with the ‘good ozone’ is located in the

35
2 The Climate Before the Collapse?

higher regions of the earth’s atmosphere – more precisely, in the stratosphere at an


altitude of 15 to 50 km. The UV light of the sun has been converting air oxygen (O2)
into ozone (O3) there forever. The ozone layer absorbs most of the dangerous UV radia-
tion. Certain gases, such as chlorofluorocarbons (CFC) – from old refrigerators, air
conditioners and aerosol cans – decompose the ozone in the stratosphere. This is why
the ozone content in the ozone layer has decreased rapidly in recent decades. A hole in
the ozone formed, mainly over Antarctica. An increased amount of UV light now reaches
earth and is contributing to an increase in skin cancer. International agreements have
limited the use of CFC. Its use has been forbidden in new plants in Germany since 1995.
In the meantime a slow recovery has been observed in the ozone layer.
Yet the ozone in the ozone layer is itself a cause of the natural greenhouse effect.
For that reason the destruction of the ozone has actually produced a light cooling
effect.
‘Bad ozone’ close to earth, on the other hand, occurs as a reaction to nitrogen oxides,
oxygen and sunlight and causes the notorious summer smog. Ozone acts as an irritant
and in large concentrations is very toxic. In the natural ozone layer above an altitude of
15 km this is not a problem, but it is a problem in the troposphere close to earth.
Furthermore, the tropospheric ozone close to earth also contributes to the greenhouse
effect. Unlike what happens in the ozone layer, the ozone concentration close to
earth is increasing. To this extent the ozone close to earth is what is warming up
the earth.

CFC was used as a coolant in air-conditioning systems and refrigerators. This not
only contributed to the greenhouse effect but also damaged the ozone layer. As
a result, the Montreal Protocol stipulated an end to CFC production worldwide,
allowing long periods of transition. HFCs are often used today as a CFC substitute.
These substances no longer use chlorine as an ingredient. Although this means
that they can no longer damage the ozone layer, HFC is still a problem for the
greenhouse effect. The HFC cooling agent R404A has a specific greenhouse poten-
tial of 3260. Therefore, 1 kg of R404A is just as harmful as 3.26 tons of carbon
dioxide. But there are substitute substances available that are not damaging to
the climate. Unfortunately, these are currently only being used in a limited number
of areas.
Scientists have now developed a relatively good understanding of the different
effects of individual greenhouse gases and other factors on the earth’s warming.
The greenhouse gases that result from the activities of people increase the reflection
in the atmosphere and thus cause an increase in global warming. Figure 2.9 shows
the contribution of the different gases. The rise in solar activity is also causing a
slight increase in radiation. As described earlier, aerosols released through the
activities of people and the resulting cloud formations provide a cooling effect. Even
the ozone layer and changes in land use are producing a light cooling. If one jux-
taposes the cooling and the warming effects, it is the warming that predominates.
An increase in greenhouse gases and a decrease in aerosols through further improve-
ments in air purification could even cause a considerable increase in warming over
the next few years.

36
2.3 Outlook and Recommendations – What Lies Ahead?

anthropogenic natural cooling


greenhouse effect effects effects
3.5
water vapour
3
laughing gas (N 2O)
near-ground ozone (O3)
2.5
CFC, HFC and halons
radiative forcing in W/m²

2
methane CH4
1.5
carbon dioxide CO2
1

0.5 increased
solar activity
0

-0.5 aerosols (dust, smoke)


and cloud f ormation
-1
change in land use
ozone hole
-1.5

Figure 2.9 Causes of global warming. Data: IPCC (Intergovernmental Panel on Climate
Change, 2007), www.ipcc.ch.

2.3 Outlook and Recommendations –


What Lies Ahead?

No one can predict how many greenhouse gases will be emitted in the future.
Climate models can, however, indicate the effects of different emission levels. For
this purpose climate researchers have designed different models that describe the
effects of changes in greenhouse gas concentrations. On the basis of these models,
they developed complex computer programmes. To check a computer model, one
first tries to reconstruct the changes of the past. So far this has been a successful
approach and future projections have achieved a high quality.
The Intergovernmental Panel on Climate Change (IPCC) regularly publishes reports
on climate change and studies the effects of climate. The environmental programme
of the United Nations (UNEP) and the World Organization for Meteorology (WMO)
initiated the IPCC in 1988. The most competent climate researchers in the world
are involved in the preparation of the IPCC reports and, as a result, this group has
a good reputation internationally.

■ www.ipcc.ch Intergovernmental Panel on Climate Change


■ www.unfccc.de United Nations Framework Convention on Climate Change

It is virtually impossible to predict parameters such as population growth or the


point when climate protection measures will be implemented. Therefore, the

37
2 The Climate Before the Collapse?

calculations in the models are restricted to certain scenarios. These indicate what
can be expected in the best case and the worst case scenarios. The conclusions of
the climate researchers show that there is still time for us to have a drastic influence
on future developments.
What is clear is that there is no way to stop climate change completely. However,
determination and appropriate countermeasures can keep the consequences of
climate change within a manageable limit and largely save the climate. If we do not
change course during the next ten years, climate change will very quickly reach a
point of no return with no hope of stopping it. Table 2.3 shows how we can still set

Table 2.3 Consequences of climate change – the climate can still be saved.

This is how the climate could This is what threatens us in the worst case
develop if it were protected scenario
worldwide
■ Greenhouse gas transmissions ■ All fossil energy sources are almost used up
are reduced worldwide by at and worldwide greenhouse gas emissions
least 50% by 2050 and by show a definite further increase.
almost 100% by 2100.
■ The average global temperature ■ The increase in average global temperature
increase remains below 2 °C. reaches values of 5 °C and higher.
■ Sea levels rise by 20 to 30 cm ■ Sea levels rise by a metre or more by 2100.
by 2100.
■ The ice in Greenland remains ■ The entire ice mass in Greenland melts
intact. away, directly resulting in a 7 m rise in sea
levels alone in the long term.
■ In the long term the rise in sea ■ In the long term the sea level rises by more
level remains around a metre or than 30 m.
less.
■ Coastal regions and cities are ■ Entire coastal regions and cities such as
saved through slight Hamburg, Rotterdam and Miami sink into the
improvements in coastal sea. Even cities on higher ground could be
protection measures. threatened in centuries to come.
■ The Gulf Stream weakens only ■ The Gulf Stream fails completely. The
slightly. climate conditions in Europe undergo
extreme changes.
■ Heatwaves, periods of drought ■ Heatwaves, periods of drought and extreme
and extreme precipitation precipitation increase dramatically and in
increase. some regions of the world threaten the
foundations of life.
■ Migrations of populations are ■ The consequences of climate change trigger
limited locally due to climate a major migration of people who flee their
change. homes due to climatic conditions. Global
tension rises dramatically as a result.

38
2.3 Outlook and Recommendations – What Lies Ahead?

the course today. In the worst case scenario, the entire mass of Arctic ice will melt
away. The melting of the ice mass in Greenland alone could cause the sea level to
rise 7 m over the next several centuries. Figure 2.10 shows what the effect would
be on Germany.

North Sea
Baltic Sea

100 km

Figure 2.10 Threatened areas in Northern Germany if the sea level were to rise by 7 m in
the long term. Graphics: Geuder, DLR.

2.3.1 Will It be Bitterly Cold in Europe?


The American climate researcher Wallace Broecker wrote an article in 1987 titled
‘Unpleasant Surprises in the Greenhouse’. The article, which caused quite a sensa-
tion at the time, warned about potential surprises from the greenhouse effect that
could affect us very seriously: global warming could, paradoxically, turn Europe
bitterly cold.
While studying the history of the climate from ice drill cores in the Greenland ice
mass, climate researchers determined that the world’s present climate is quite stable
– but it has only been this way for about 10 000 years. The average yearly tempera-
ture on earth has never deviated more than one degree from the long-time average
during this time. However, if one were to look back another 2000 years, one would
find that major jumps occurred in temperatures even during short periods of only a
few years. Until now it was always thought that temperature changes as well as the
transition from ice ages to warming periods only happened very slowly. However,
this opinion had to be corrected as a result of the drill core studies.
Today it is assumed that the climate tends to behave erratically on its own, thus
making it a powerfully non-linear system (Rahmstorf, 1999). Linear systems are
easy to understand. The stronger that the attraction is, the clearer the reaction is.
For example, a water tap approximates a linear system. With two turns the amount
of water that flows is double the amount of one turn. Climate is a complex non-linear
system. It has the tendency of self-regulation and suddenly jumps into a different

39
2 The Climate Before the Collapse?

state when a critical point has been reached. The human body is another example
of this. On its own the human body is able to keep its body temperature quite con-
stant at 37 °C, even during major changes in ambient temperature. However, if a
person becomes seriously hypothermic, the body temperature quickly drops at a
certain point.
The Sahara is an example of very erratic climate change. Several thousand years
ago the Sahara was green. The vegetation there absorbed moist air from the Atlantic
and monsoon rains. Due to very minor changes in the earth’s orbit and in the incli-
nation of the earth’s axis, the conditions for the monsoon rains deteriorated over a
period of thousands of years. The vegetation managed to survive until a certain point
around 5500 years ago. But then within a short period the vegetation died and the
Sahara turned into a dry desert. A new stable state then developed, which continues
until today. The populations of large areas had to flee due to the dryness of the
terrain. The Nile valley in Egypt became a place of refuge for the environmental
refugees of the day. This contributed towards the creation of the Pharaonic high
culture and was perhaps even the basis for the biblical stories that have been passed
down through the ages.

■ http://www.pik-potsdam.de/%7Estefan Internet page of climate researcher


Stefan Rahmstorf at the Potsdam
Institute for Climate Impact
Research

The origins of the erratic climate changes that occurred many years ago are sus-
pected to lie in changes in ocean currents, which continue to have a major effect on
our climate even today. For example, Berlin and London are located further north
than Quebec in Canada. A comparison of temperatures shows that Western Europe
is about five degrees warmer than it should be considering its latitude. The reason
for this is a gigantic self-regulating heat transport machine: the Gulf Stream.
In Central America the sun heats the water masses of the Caribbean Sea and the
Atlantic. The Antilles Current and the currents from the Caribbean merge together
to become the Gulf Stream, which is 50 km wide and transports enormous masses
of warm saltwater northwards along the American coast. These currents are so large
they can be seen from space. From North America the current then moves straight
across the Atlantic, continues along the European coast to Norway and ends in the
European part of the North Sea (Figure 2.11). Some of the heat is absorbed by the
air beforehand, and this air then moves with westerly winds towards Europe. This
explains why winds from the west always promise relatively mild temperatures to
some parts of Europe.
In the North Sea between Norway, Iceland and Greenland the warm saltwater comes
together with colder, less salty water and cools off quickly. This water increases in
density, becomes heavier and consequently sinks to the bottom quickly. In the North
Sea 17 million cubic metres of water sink to a depth of 3000-4000 m per second.
This corresponds to around 20 times the entire capacity of all rivers on earth. The

40
2.3 Outlook and Recommendations – What Lies Ahead?

water masses then return to Central America as cold deep water flow, to be heated
up again there.
Due to climate warming, large quantities of meltwater (melted snow and ice) pour
into the North Sea. This fresh water dilutes the warm saltwater. It has already been
determined that the salt content has fallen considerably. If the salt content decreases
even further, the weight of the water will not be sufficient in the future to allow it
to sink down. The enormous Gulf Stream water pump would get out of step and
very suddenly come to a halt.

nt
Curre
Canary

Antilles Curre
nt
Ca
rib
be
an
Cu
rre
nt

Figure 2.11 Principle of the


Gulf Stream.

Climate researchers suspect this was precisely the cause of the temperature jumps
11 000 to 12 000 years ago. After the ice age enormous amounts of meltwater from
the melting glaciers poured through the St Lawrence River into the North Atlantic.
It diluted the sea water with fresh water, so that the uppermost layer of sea water
became noticeably less salty and the water lighter. Therefore, the water no longer
sank to the depths, despite strong cooling. The Gulf Stream was ‘turned off ’ and
for many years large areas of Northern Europe and Canada were exposed to extreme
cold and covered in ice. As a result, the amount of meltwater decreased again. The
salt content of the water rose. The warm water heater, the Gulf Stream, started
working again.
This is exactly the fate that we are facing again today. No scientist can accurately
predict just how much global warming must increase before the Gulf Stream stops
working. But the assumption is that the critical point will have been reached once
worldwide warming increases by 3 °C. We already have increased the temperature
by a good 0.7 °C. The consequences for us would be catastrophic. The climate would
be erratic for years. Ultimately, it would be so cold and dry in Europe that any kind

41
2 The Climate Before the Collapse?

of farming would be impossible. Yet, according to the calculations of the climate


researchers, this new climate state would then remain stable again for thousands of
years.

2.3.2 Recommendations for Effective Climate Protection


The only thing that could help to keep climatic effects at a manageable level and
prevent some nasty surprises, such as the disappearance of the Gulf Stream, would
be a speedy reduction in greenhouse gas emissions. By 2050 greenhouse gas emis-
sions worldwide would have to be reduced to half of what they were in 1990.
The industrialized countries are currently causing the biggest per-head output of
greenhouse gases. The developing and emerging countries justifiably want to catch
up in terms of prosperity and energy demand. Therefore, the industrialized countries
have to make the biggest contribution towards reducing greenhouse gas emissions.
Emission reduction targets were formulated for countries such as Germany, Great
Britain and the USA as far back as the late 1980s. Today the need to meet these
targets is even more critical (Enquete-Kommission ‘Vorsorge zum Schutz der
Erdatmosphäre’ des 11. Deutschen Bundestages, 1990). Based on decisions made
in 1990, the following targets have been set as a minimum goal for industrialized
countries:
■ Reduction in greenhouse gas emissions of 25 to 30% by 2005
■ Reduction in greenhouse gas emissions of 40 to 50% by 2020
■ Reduction in greenhouse gas emissions of 80% by 2050
■ Reduction in greenhouse gas emissions of 90% by 2100

2.4 Difficult Birth – Politics and Climate Change

2.4.1 German Climate Policy


The realization that the consequences of an increase in climate warming represent
a previously unknown type of threat for mankind finally struck home with the politi-
cians in the 1980s. At the time the public was developing a keen awareness of the
environment. Dying forests, respiratory illnesses due to air pollution and the risk of
nuclear power were topics that had a big impact on people. It was no longer possible
to ignore global environmental problems.
In 1987 the German Bundestag (Lower House of Parliament) appointed the cross-
party Enquete commission to deal with the issue of protecting the earth’s atmos-
phere. The commission studied the problems relating to climate change and drew
up clear recommendations for reducing greenhouse gases (see above).
The first target of a 25% reduction by the year 2005 was adopted by the conserva-
tive government of Helmut Kohl. After reunification the policy was extended from

42
2.4 Difficult Birth – Politics and Climate Change

West Germany to the unified Germany. As we see today, this was a clever move.
According to the statistics, there was a 17% reduction in CO2 emissions by 2005.
As a result, Germany likes to present itself as a forerunner in climate protection on
the international stage. However, most of the CO2 savings can be attributed to the
effects of reunification. A major part of the industry in the former East Germany,
with its high level of energy consumption and carbon dioxide emissions, collapsed.
The output of greenhouse gases was thereby reduced to around half of what it had
been, whereas in West Germany nothing much was changing in terms of climate
protection.
After the change in government in 1998, the ‘Red-Green’ government, a coalition
between the Green Party and Social Democrats, also adopted the climate protection
targets for the year 2050. However, the measures introduced were not adequate for
an effective climate protection policy that would make the targets realistic. The first
target of a 25% reduction by 2005 was missed by a long shot (Figure 2.12). Even
the current German government is not moving ahead quickly enough to reduce
greenhouse gas emissions.

reference year 1990 goal -25 % goal-50 %


1200
Mt
carbon dioxide emissions _

1000 total Germany

Kyoto goal
800 trend

Western Germany trend


600

goal
400
goal

200 goal
Eastern Germany
trend
0
1985 1990 1995 2000 2005 2010 2015 2020

Figure 2.12 Energy and process-related carbon dioxide emissions in Germany.

On the international stage Germany had already qualified its climate protection
targets in 1997. Under Angela Merkel, the environmental minister at the time, the
government’s own target was a reduction of 25% by 2005. Within the framework
of the Kyoto Protocol, a reduction target for greenhouse gases of only 21% was
promised by the year 2012. In addition to carbon dioxide, the calculations of the
Kyoto protocol take into account other climate gases, such as methane, nitrous
oxide, HFC and CFCs. This clearly improves Germany’s chances of at least reach-
ing the Kyoto target. It has had great success particularly in the reduction of methane
emissions – for example, through recycling.

43
2 The Climate Before the Collapse?

The United Kingdom was the only Western industrialised country to achieve a
significant reduction in greenhouse gas emissions of about 15% between 1990 and
2005. Carbon dioxide emissions went down by 6% due to the change from coal to
natural gas in some parts of the energy supply.

2.4.2 International Climate Policy


In 1979 the main theme at the first UN conference on world climate in Geneva was
the threat to the earth’s atmosphere. However, many special-interest representatives
had their doubts about the essence of the threat. So it was initially agreed that further
research would be needed before any particular measures could be considered.
In Rio de Janeiro in 1992 the members concluded for the first time that measures
to protect the environment were now absolutely essential. The first targets for reduc-
ing greenhouse gas emissions were then agreed in 1997 in what was referred to as
the Kyoto Protocol. Although these targets still do not go far enough, they have
succeeded in putting international pressure on countries to implement the measures
necessary to protect the climate.

The Kyoto Protocol


The first world summit on climate, the UN Conference for Environment and
Development (UNCED), took place in Rio de Janeiro in 1992. One result of
the conference was the Climate Framework Convention on Climate Change (UNFCCC),
where the signatories made a commitment towards preventing dangerous anthropogenic
disruption to the climate system, slowing down global warming and lessening its
consequences.
This rather vague agreement was followed by the Kyoto Protocol, which was drawn up
in 1997 at the third conference of the treaty states of the UNFCCC in the Japanese city
of Kyoto. It obligates all the treaty parties listed in Figure 2.13 to reduce their emissions
of greenhouse gases by an average of 5.2% by 2012 at the latest. However, different
restrictions apply to individual states. The protocol actually allows certain countries,
such as Spain, Portugal, Greece and Australia, to increase their emissions. The EU states
should be able to achieve their targets together. This means that low-emitting EU states
can compensate for the high emissions of other EU states.
In addition to carbon dioxide (CO2), the following greenhouse gases are also being
considered and converted into carbon dioxide equivalents: methane (CH4), nitrous oxide
(laughing gas, N2O), partially halogenated chlorofluorocarbons (CFC), perfluorocarbons
(PFC) and sulphur hexafluoride (SF6). Measures initiated in other countries can also be
counted as reductions. The Kyoto Protocol plans no restrictions for developing and
emerging countries. The Kyoto Protocol came into effect in 2005. The USA is the only
country that has not yet ratified it.

Most industrialized nations have so far shown meagre results in their tackling of
climate warming. The only countries that have recorded definite reductions in
greenhouse gas emissions are those that were formerly in the Warsaw Pact. The
reason for this is due more to the economic upheavals of the 1990s than to a con-

44
2.4 Difficult Birth – Politics and Climate Change

sistent policy on climate protection. In the meantime CO2 emissions have been rising
again in these countries too.
In most Western countries emissions are increasing at the same levels as before
(Figure 2.13). Spain is the worst culprit for this. Because its emissions in 1990,
compared to countries like Germany and the USA, were still comparatively low,
Spain was actually allowed an increase of 15%. However, by 2004 Spain’s actual
emissions had already risen by 51%. Only Germany and Great Britain have really
succeeded in achieving significant reductions, although the bulk of the reductions
in Germany are attributed to the upheavals in the former East Germany.

60 %
changes between 1990 and 2006
50 %
goals of the Kyoto protocol
40 %
30 %
20 %
10 %
0%
-10 %
-20 %
-30 %
-40 %
-50 %
-60 %
-70 %
Luxem bourg

Liechtenstein
Estonia

Netherlands
Rom ania

Sweden
Russia

Czech Rep.
Lithuania

Slovenia
Hungary

Germ any

Denm ark

Iceland
Monaco

Norway

Portugal
Greece
Latvia

Austria
France

Ireland
Slovakia
Belarus

Spain
UK

Finland

Canada

Australia
USA
Croatia
Bulgaria

Switzerland
Poland

Italy
Belgium

New Zealand
Japan
Ukraine

Figure 2.13 Changes in greenhouse gas emissions with no change in land use between 1990 und
2004 and targets from the Kyoto Protocol. Data: UNFCCC (United Nations Framework Convention on Climate
Change, 2008).

The USA, which essentially sees no chance of achieving its targets for climate
protection during the next few years, did not even ratify the Kyoto Protocol.
Nevertheless, the very different major changes that individual countries have made
shows just how much can be done to protect the climate. If these plans are taken
seriously, a reduction in greenhouse gases of 50 to 80% by the year 2050 should
be easily achievable.

45
2 The Climate Before the Collapse?

2.5 Self-Help Climate Protection

The governments of most countries today are taking a half-hearted approach to


addressing the issue of climate protection. Certain special-interest groups would
invariably suffer financially if consistent policies were pursued in this area. Therefore,
these groups are trying to use their political influence to delay any serious imple-
mentation of climate protection measures.
However, this is not a reason to fall into a state of lethargy. On the contrary:
Everyone has the chance to make a contribution towards protecting the climate.
Savings of 50, 80 or even 100% in greenhouse gas transmissions are easily achiev-
able right now. If more and more people follow a good example, the results of their
efforts will help to protect the climate. But a single measure on its own is not enough
to stop global warming. Instead it will take a combination of many different kinds
of measures. Renewable energies in particular can play a major role here, as this
book will show.
Active climate protection does not demand any major sacrifices. If everyone tried
to implement the following measures, it would be enough to save the climate:
■ Avoid buying products made of tropical woods that do not come from sus-
tainable sources. The FSC stamp can help in a purchase decision.
■ Avoid equipment and appliances (refrigerators, air conditioning units,
heating pumps) containing coolants with HFCs because they are responsible
for more than 10% of the greenhouse effect. If this equipment is not available
without HFCs (e.g. car air conditioning systems), the buyer should put pres-
sure on the manufacturer.
■ Be consistent in recycling, reduce the use of animal feed and increase the
use of organic and local products.
■ Be consistent in saving energy in all areas (See Chapter 3).
■ Build and finance new renewable energy systems and installations (see
Chapters 5 to 14).

46
3
3 From Wasting Energy to
Saving Energy and Reducing
Carbon Dioxide
In modern industrial societies energy is still available at relatively moderate prices
and without any restrictions. Whenever we want it, we can consume as much energy
as we like. However, the general increase in energy prices in recent years has
brought about a new awareness of the value of energy. High energy consumption
is making more and more of an impact on our finances. Yet the equation is very
simple: saving energy also means saving money. But the concept of saving energy
is not totally new.
Until the 1970s it was accepted that real economic growth and increasing pro-
sperity would also require the use of more energy. The concept of saving energy
developed when the oil crisis in the 1970s led to an explosion in oil prices and put
a brake on growth. Numerous tips, appeals and stickers in the 1980s were aimed
at encouraging citizens to save energy. In any event, the trend towards stopping
energy waste was successful at the time. As prices started dropping again in the
1990s, the original goal was largely ignored and energy was again thoughtlessly
wasted.

3.1 Less Efficient – Energy Use and Waste Today

High energy consumption is not a necessary prerequisite for maintaining our pros-
perity and living standard. In fact, energy use is coupled with enormous losses.
Around 35% of the primary energy used is already lost in the energy sector through
power plant waste heat or during transport even before it reaches the consumer.
Various devices and machines then produce the desired usefulness, such as light,
heat, and driving power for machines and vehicles. This too leads to high losses.

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

47
3 From Wasting Energy to Saving Energy and Reducing Carbon Dioxide

Light bulbs and vehicle combustion engines are particularly inefficient and trans-
form around 80% of energy used into undesirable waste heat. Even ‘useful’ energy
is often wasted, for example when lights are left on in empty rooms, poorly insulated
buildings are heated or people drive round and round the block to find a parking
space. If one looks at all the losses incurred, it emerges that at best only 20% of
original primary energy is being used sensibly (Figure 3.1).

Figure 3.1 During transport or conversion around 80% of energy in Europe is lost or not
used sensibly.

When it comes to certain aspects of the lifestyles of industrialized societies, the


percentage of sensible energy use is even much smaller. However, the sort of person
who tries to talk friends out of taking long flights or buying a new car with all the
extras, or keeps telling his family to turn down the central heating, tends not to be
very popular. The choice of one’s own lifestyle, finances permitting, is considered
one of those individual freedoms that no one else has a right to influence. Therefore,
saving the environment does not mean looking for scapegoats but instead seeking
solutions that meet both lifestyle and environmental needs.
This does not mean that everyone can abuse his or her right to a certain lifestyle
and thoughtlessly treat energy and the environment with disrespect. People who
against their better judgement refuse to accept a technology that would enable them
to retain their lifestyle but at the same time considerably reduce energy consumption
and save the environment are acting irresponsibly. This also applies to policies that
do not strive towards a speedy introduction of the most optimal technologies.
Many small steps that could help create an environmentally compatible society are
also not being implemented due to a lack of knowledge or appropriate awareness.
Many of the problems related to energy use are extremely complex. Optimal solu-
tions to these problems often depend on a number of different factors.
For example, one issue that is often debated is whether a gas or an electric stove is
more efficient. Compared to an electric stove, a gas stove produces more waste heat.
Anyone who has ever scorched a potholder on a gas hob will be able to confirm
this. Yet gas has the reputation of being the ecologically better alternative.
Conventional coal-fired, natural gas and nuclear power plants for electricity genera-

48
3.1 Less Efficient – Energy Use and Waste Today

tion do not work efficiently. More than 60% of the primary energy used in these
facilities is lost as power plant waste heat (Figure 3.2). If the electricity used to boil
a litre of water comes from a coal-fired power plant, it is releasing 156 g of carbon
dioxide. In contrast, burning natural gas on a gas stove releases about 56 g of carbon
dioxide. However, the losses that occur when gas is transported from the place of
extraction to the end consumer are also a problem. Natural gas essentially consists
of methane, which is considerably more damaging to the environment than carbon
dioxide. Therefore, even the loss of a few grams can be harmful. If transport losses
are at the rate of about 10%, a gas stove is still causing a lower rate of greenhouse
gas emissions than an electric stove using coal-generated electricity. However, the
pipelines in the areas where natural gas is exploited are sometimes in very poor
condition. The use of natural gas from these sources can completely cancel out the
advantages of gas stoves.

Figure 3.2 Energy and environmental effects from boiling water using an electric versus
gas stove.

A gas stove should normally be a better alternative than an electric one – as long
as the energy used is not supplied by fossil power plants. ‘Green’ electricity sup-
pliers offer carbon-free electricity from renewable energy plants. Using this kind of
electricity may make an electric stove more environmently friendly than a gas stove.
In Norway all electricity is supplied by renewable energy plants. In this case, electric
stoves are generally preferred. Finally, the use of an electric kettle to boil water is

49
3 From Wasting Energy to Saving Energy and Reducing Carbon Dioxide

an efficient way to save energy and reduce carbon dioxide and puts the user ahead
of the game. Anyone who worries generally about saving energy should first look
at the areas that use the most energy. What will come as a surprise to many is that
private households in Germany use more secondary energy than industry and the
transport sector (see Figure 3.3). This applies to many countries in the world.

traffic
households 28.7 %
28.8 %

Figure 3.3 Percentage of different


trade and sectors in secondary energy
services
15.7 % industry consumption in Germany. Data:
26.8 % (Tscheutschler, Nickel and Wernicke, 2007).

This was not always the case. In 1990 industry was still the biggest consumer of
secondary energy. Yet while industry reduced its secondary energy use by over 17%
between 1990 and 2005, private households and the traffic sector increased their
consumption by at least 10% during the same period.
The main cause of this increase in domestic energy consumption is the boom in
sales of electrical devices and appliances. This increase is particularly noticeable in
the areas of communication and entertainment electronics. Modern widescreen
plasma televisions can easily use twice as much energy as an old cathode ray tube
TV set. But there are also other hidden things that eat up energy. As the possibilities
for saving energy in the household and transportation are particularly easy to imple-
ment, these options head the list of any analysis on energy savings.

3.2 Personal Energy Needs – Easily Saved at Home

3.2.1 Domestic Electricity – Money Wasted


Europeans take the supply of electric energy so much for granted that we find it
hard to imagine doing without electricity even for a short time. Televisions, tele-
phones, computers, lights, refrigerators, washing machines and even heating do not
function without electricity. It is hard to grasp, then, that around two billion people,
one-third of the world’s population, have no access to electric power.

50
3.2 Personal Energy Needs – Easily Saved at Home

In Germany, an average three-person household consumes around 3900 kilowatt


hours of electricity at a cost of about 750 euros per year. More than 10% of the
carbon dioxide emissions from energy use in Germany come from the electricity
needs of private households (Umweltbundesamt, 2007). Furthermore, inefficient
electrical equipment accounts for a considerable percentage of electricity consump-
tion. Anyone looking for ways to cut back on consumption can quickly find ways
to save up to 30% of their outlay on energy, which equates to more than 200 euros
a year. Therefore, protecting the climate can also pay off financially. An average
US household consumes even more than 10 000 kilowatt hours per year. Hence, the
saving potential is much higher there.

Standby Losses – Power Destruction Par Excellence


Many electrical devices work at low voltages. A transformer transforms
down the mains voltage for this purpose. Most devices have a switchoff
mechanism, but it usually only switches off part of the electronics in the lower voltage
area. A transformer and often also major parts of the device electronics remain connected
to the grid and continuously consume electricity, even when in a switched-off state. In
this state they cause standby or open-circuit losses. As a result, over the course of a year
a considerable amount of power can be consumed even though the appliance is not used
very often. Power wasters like this can actually be identified using an energy consump-
tion measuring device.
There are very few cases where, for technical reasons, a device should not be completely
unplugged from the mains. Mains switches are a few cents more expensive than switches
for the lower voltage area. When large numbers of devices are involved, a manufacturer
can easily save thousands of euros without fearing a loss of sales by installing the
cheaper switches. Open-circuit loss is currently not a major consideration in purchase
decisions.

Computers and communication and entertainment electronics currently account for


around one-fifth of electricity consumption in Germany and one-tenth in the USA.
This kind of equipment often wastes obscene amounts of electricity. Many devices
have high standby consumption, which means they consume electricity even when
turned off. A device with standby consumption of only 5 watts uses a total of 43.8
kilowatt hours and costs 8 euros for electricity per year – just for the period when
it is switched off and not being used at all. Some, usually older, devices even have
standby losses of over 30 watts. On average each household in Germany wastes
around 85 euros per year due to standby losses. The German federal environment
office estimates that open-circuit losses in Germany cost about 4 billion euros per
year and produce 14 million tons of carbon dioxide. This is more than seven times
the amount of carbon dioxide emitted through energy use by the 20 million inhabit-
ants of Mozambique.
This is something that is very easy to remedy. By using a switchable multipoint
plug for their computer, television and stereo system, consumers can reduce their
open-circuit losses in a switched-off state to zero. When buying new devices

51
3 From Wasting Energy to Saving Energy and Reducing Carbon Dioxide

or equipment, consumers should ask questions about open-circuit consumption


values so that more pressure is placed on manufacturers to fit power-saving
switches.
Around 10% of the electricity used in private households in Germany is for lighting
(see Figure 3.4). In the USA the percentage is about the same. Here too a great deal
of energy, and money, is wasted. Due to inertia and unfounded prejudice against
energy-saving bulbs, millions of conventional electric light bulbs are still being
used. This is not only damaging the environment but also hurting the pocket.
Modern low-energy light bulbs produce the same amount of brightness using 80%
less electricity. An 11-watt low-energy bulb saves about one cent an hour. If this
kind of light bulb burns for two hours each day, the savings amount to seven euros
a year. Over its lifetime a low-energy light bulb can save 100 euros and 300 kg of
carbon dioxide. Therefore, the higher purchase cost is rapidly recouped. Good-
quality low-energy light bulbs also tolerate frequent on and off switching well. Tiny,
dimmable light bulbs open up possibilities to all kinds of uses (Figure 3.5). A con-
sistent approach to protecting the environment should include a gradual replacement
of all light bulbs with low-energy bulbs. However, there is a growing trend to fit
halogen lights. In terms of energy consumption, these lights are only marginally
better than conventional bulbs and, consequently, not only expensive to buy but also
to use. Energy-saving LED lights continue to be improved and will offer a viable
alternative in the future.

Figure 3.4 Low-energy light bulbs


save energy, carbon dioxide and cash
and are now available in almost all
sizes and styles. Photos: MEGAMAN,
www.megaman.de.

52
3.2 Personal Energy Needs – Easily Saved at Home

Figure 3.5 Tiny, dimmable light bulbs open up possibilities to all kinds of uses.

Electric hot water boilers are not only inefficient but also expensive. Anyone who
has an electric boiler should switch it off when leaving the house for long periods.
A considerable amount of energy can also be saved with large electrical appliances,
such as washing machines, dishwashers, refrigerators and freezers. Efficiency
should be an important criterion, especially in the selection of new appliances.
Energy efficiency-class labels and the consumption values included with these labels
provide an indication of the economy of use with different appliances.
Unfortunately, the energy-efficiency classes indicated are not always that clear at
first glance. The washing machines, dryers and dishwashers with the highest effi-
ciency are those with an A rating. With refrigerators and freezers the efficiency scale
has been extended to include classes A+ and A++. The most efficient appliances in
this category therefore have an A++ label. A person who chooses a refrigerator with
a class A++ rating over a reasonably priced appliance in energy-efficiency class B
can save 30 euros in electricity costs per year. With a 10-year lifetime for the appli-
ance, this adds up to 300 euros. The higher price of efficient devices and appliances
often pays for itself over the lifetime of the item. And the environmental benefits
are greater. Compared to one in class B, an appliance in class A++ can save up to
a ton of carbon dioxide.
If someone is not aware of how much energy a device or appliance is consuming,
he or she can use an energy-saving monitor (available for hire or purchase) to test
which appliances are energy wasters. It is advisable to take particularly inefficient
appliances out of service as quickly as possible.
A savings potential also exists in the way household goods are used. If a refrigerator
is standing right next to a stove, it has to work harder because of the stove’s heat.
Refrigerators with poor ventilation and iced-up freezer compartments also use a
large amount of extra energy. A lid on a pot when cooking can save an appreciable
amount of energy, as can a pressure cooker. Selecting the lowest temperature

53
3 From Wasting Energy to Saving Energy and Reducing Carbon Dioxide

possible on a washing machine or a dishwasher saves energy and money. With


tumble dryers the laundry should first be spun at the highest spin speed available.
But the good old clothesline is the best option for saving energy.
The following energy-saving tips summarize the key points discussed:
■ Track energy guzzlers using an energy consumption measuring device.
■ Switch off unnecessary electrical devices and lights.
■ Switch off all devices with standby mode using a switchable multipoint
switch.
■ Replace light bulbs and halogen bulbs with low-energy bulbs (compact fluo-
rescent lights) or LED lights.
■ When buying electrical goods, pay attention to their energy consumption
ratings.
■ Always buy the most energy-saving household goods (energy efficiency
class A, or A++ for refrigerators and freezers).
■ Do not place freezers next to items that produce heat (ovens, radiators).
■ Try to thaw frozen goods in a refrigerator.
■ Defrost refrigerators and freezers regularly.
■ Only run washing machines when they are full and operate them at the lowest
possible temperature. When tumble drying clothes, use a high spin speed
during the washing cycle.
■ When cooking, use lids on pots and frying pans, or use pressure cookers.

3.2.2 Heat – Surviving the Winter with Almost No Heating


In many countries the lion’s share of secondary energy consumption in private house-
holds goes for heating. Around three-quarters of secondary energy in households is
used for this purpose. However, reducing heat in a home does not necessarily mean
having to cope with frosty temperatures. With good insulation and modern building
technology pleasant room temperatures can be achieved with energy savings of up
to 90%. In other words, ten energy-efficient buildings can be kept warm with the
same amount of energy it takes to heat the average poorly insulated old house. At the
same time the carbon dioxide emissions and heating costs drop to one-tenth.
Many people live in rented accommodation. This can cause a dilemma, because
energy-saving measures are usually linked to investment. This is a cost that the
landlord first has to bear, with the renters becoming the beneficiaries – and because
they didn’t pay for the measures in the first place, they don’t have any particular
incentive to implement them. However, from the standpoint of saving energy, even
actual homeowners lag behind when it comes to making use of the options that are
available.
Yet not all energy-saving measures cost money. The following changes to heating
behaviour can save a considerable amount of heat energy and, consequently, reduce
heating costs:

54
3.2 Personal Energy Needs – Easily Saved at Home

■ Do not select a room temperature that is higher than necessary. Each extra
degree consumes around 6% more heat energy.
■ Turn down the heat at night and when no one is at home.
■ Close blinds, shutters and curtains at night.
■ Avoid leaving windows open all the time to air rooms; instead briefly open
windows completely a few times a day to bring in fresh air.
■ Do not enclose, block or hide radiators behind curtains.
Even minor investments can help reduce energy use considerably. Appropriate
measures can include draught-proofing windows and doors, and upgrading thermo-
stat valves.
In terms of saving energy, a tremendous amount can be achieved in new-build
homes or planned renovations. In the case of three-litre and passive houses, heating
requirements can be slashed to as little as one-tenth of what is normally needed in
conventional houses. The heating required in an average apartment block in Central
Europe is around 150 to 200 kilowatt hours per square metre of living area per year
(kWh/(m2 a)). Old buildings with poor insulation can even show usage values of
over 300 kWh/(m2 a). A new-build still needs around 100 kWh/(m2 a). In contrast, a
three-litre house only needs 30 kWh/(m2 a), whereas a passive house gets by with
even less than 15 kWh/(m2 a) (Figure 3.6). These figures can vary for other climate
regions. However, in warmer countries there is a higher demand for cooling than
for heating. The energy-saving possibilities with optimal insulation are similar.

Figure 3.6 Comparison of heating requirements and heat loss in houses with different
insulation standards in kilowatt hours per square metre of living area per year in Central
European climates (kWh/(m2 a)).

A Low-Energy House is Not Necessarily a Low-Energy House


Anyone interested in building a low-energy house will first have to grapple
with a variety of different terms and concepts that sometimes confuse even
the experts. The following should help to unravel some of this confusion.
Low-energy house – These are especially well insulated buildings that have low heating
requirements. Most countries have not legally specified this concept. Normal new-builds
are often unjustifiably labelled as being low-energy or energy-saving buildings.

55
3 From Wasting Energy to Saving Energy and Reducing Carbon Dioxide

Three-litre house – A three-litre house colloquially is a house that requires around 3


litres of heating oil per square metre of living area per year (approx. 30 kWh) in Central
European climates. This corresponds to a primary energy requirement of around 60 kWh/
(m2 a), which also includes the efficiency of the heating system and hot water boiler.
Passive house – A passive house can manage almost without any heating at all. The
annual heating requirement is less than 1.5 litres of heating oil (approx. 15 kWh) per
square metre of living area. This equates to a primary energy requirement of less than
40 kWh/(m2 a).

Building insulation and window type have an important effect on heating require-
ments (Figure 3.7). The U-value is a comparison value for the quality of heat insula-
tion. This value indicates the heat loss per square metre of wall and window space
and per degree of temperature difference between indoors and outdoors. Low
U-values therefore mean low heat loss.

Figure 3.7 Effect of window type and insulation on heat loss.

Compared to non-insulated outer walls, walls with high-value heat insulation can
achieve a reduction of the U-value by more than a factor of 10 in an optimal case.
Conventional insulation against heat loss of around 20 cm would be required. This
is easily achievable with prefabricated houses using wood frame construction.
Vacuum-insulated materials with a thickness of around 2 cm that provides the same
insulation results have recently been developed. With this type of insulation a core
material of pyrogenic silicic acid is packed and evacuated into an airtight foil. It is
important that the special foil of the vacuum heat insulation prevents any air from
seeping in over long periods of time.

56
3.2 Personal Energy Needs – Easily Saved at Home

Along with the insulation of the walls, the structure of the windows is also an
important element. Compared to conventional thermal double glazing, triple glazing
can reduce heat loss from windows by over 40%. Vacuum glazing is a new technol-
ogy on the market in which air is evacuated from the gap in the double glazing to
create a vacuum. A perfectly sealed edge prevents any penetration of air. Small
spacers between the layers of glass provide the necessary stability so that the exterior
air pressure does not crush the two glass panels together. Spacers made of glass are
hardly noticeable.
In addition to the type of glazing selected, the type of window frame is also impor-
tant. The frame of a window can cause further heat loss. Therefore, for the U-value
a differentiation is made between a UG value (g for glass) for the actual glass panel
and the lesser UW value (w for window) for the complete window, including the
frame. Some manufacturers only provide a general U-value. In most cases, they
mean the UG value.
Windows are not just sources of heat loss. They also let in sunlight, which in the
winter helps to heat rooms. For an optimization of solar energy gain, south-facing
windows in colder climates in the Northern hemisphere should be as large as pos-
sible, whereas those facing north should be smaller. Exterior shutters or blinds on
the sunny side are important to prevent well-insulated buildings from becoming too
hot in the summer.
Buildings with optimal insulation are comparatively airtight. For good air quality
in the winter fresh air should be let into rooms frequently. But airing rooms also
causes considerable heat loss. This is remedied through the use of controlled build-
ing ventilation systems (Figure 3.8). Ventilators blow fresh air into living areas and

Figure 3.8 Principle of controlled ventilation with heat gain.

57
3 From Wasting Energy to Saving Energy and Reducing Carbon Dioxide

extract stale air from kitchens and bathrooms. The stream of fresh air is run from
outside through a cross-heat exchanger past the vitiated air. The stale air emits up
to 90% of its heat to the cold, fresh air from outside. So the heat remains in the
building. The same system can be used to keep a building cool when outdoor tem-
peratures are high.
Controlled ventilation is often perceived as having a negative effect on living
conditions, but the opposite is actually the case. A gentle draught from such a
system is barely noticeable. Constant optimal ventilation prevents dampness forming
in walls or mould building up. An air filter in the ventilation system will keep out
some of the pollutants in the outside air and also make it difficult for insects to get
inside. Of course, anyone who feels that it is necessary can open the windows
anyway – even if this is really no longer necessary from a practical point
of view.
An air supply system can also be combined with a ground heat exchanger. This
involves laying a pipe through the soil in the garden to supply incoming air. In
winter the soil heats the fresh air and in summer it cools it.
An ultra high energy-saving house requires an outlay of additional costs, but over
the years these costs pay for themselves as energy prices rise. In some countries,
such as Germany, banks offer particularly good low-interest loans to promote
energy-saving measures in new houses and in the renovation of old buildings. In
principle, those who want to keep heating requirements carbon-neutral have the
following options:
■ Solar thermal systems
■ Biomass heating
■ Heat pumps using electricity from renewable energies
■ Heating systems based on renewable hydrogen
Solar thermal systems usually supplement other types of heating. Later chapters in
this book discuss these different variants in detail. As the potential for certain pos-
sibilities such as biomass heating is limited in many countries, the other options for
saving energy described above should be implemented as far as possible before
renewable energies are used.

3.2.3 Transport – Getting Somewhere Using Less Energy


The transport sector is responsible for about one-fifth of carbon dioxide emissions
caused by energy use. Whereas it seems that major savings can be achieved rela-
tively quickly with electricity use and in heat generation, the transport sector is more
problematic. In recent years increased mobility and the travel bug have managed to
cancel out any savings in the fuel consumption of cars and aeroplanes. Cheap air-
lines that sometimes offer plane tickets for the price of an underground fare and the
trend towards petrol-guzzling SUVs are prime examples of this development. Added
to this are the even longer distances involved in transporting goods around the world
as a result of globalization.

58
3.2 Personal Energy Needs – Easily Saved at Home

Turning back the clock and reducing the number of cars, planes, trucks and ships
would be a hopeless undertaking. Modern economies and modern lifestyles have
become so fast-paced that they demand increased mobility. Holidays tend to be
short, so air trips are necessary not to eat into travel time. Weekend breaks by air
provide relaxation in hectic schedules. Our economic growth also is based on high
levels of exports and reasonably priced raw materials, both of which are associated
with long-distance transport.
Nevertheless, to reduce carbon dioxide emissions, we simply have to get from A to
B using less energy. In the medium term, carbon-free transport must be available
to cut out emissions completely.
Selecting the right means of transport is the best option for a quick reduction in
energy requirements and emissions in traffic. For example, per passenger kilometre
the secondary energy requirement of a train is less than one-fifth that of a car. To
compare the two modes of transport, the energy consumption of the train is divided
by the average number of passengers and the energy consumed recalculated into
petrol.
Although the secondary energy consumption of a train is still less than that of a car
by several orders of magnitude, the difference in carbon dioxide emission is not as
dramatic (Figure 3.9). This is because about one-half of a train’s electricity comes
from coal-fired power plants that in turn produce large amounts of waste heat in
generating the electricity. In Germany around 24% of a train’s electricity comes
from nuclear power stations and more than 10% from renewable energies. In coun-
tries that use a higher proportion of renewable energies in electricity generation,
such as Norway, Austria and Switzerland, the carbon dioxide emissions of the trains
are considerably lower than in Germany.

passenger passenger
4 3 2 1 person 4 3 2 1 person
car car

aeroplane aeroplane

m aglev maglev
train train

city train city train

coach coach

Intercity- Intercity-
Express Express
intercity intercity
train train

0 1 2 3 4 5 6 7 8 0 50 100 150 200 250


energy demand in litres petrol per 100 km CO2 emissions in grammes per km

Figure 3.9 Secondary energy requirements and carbon dioxide emissions per person for different
means of transport with an average use load (data for trains is estimated from the German electricity grid).

The calculations shown here are based on average values. The average carbon
dioxide emissions of new cars in Europe are currently around 160 g of carbon

59
3 From Wasting Energy to Saving Energy and Reducing Carbon Dioxide

dioxide per kilometre. Older vehicles and thirsty new cars have higher emissions,
whereas economical small cars sometimes have considerably less. Carbon dioxide
emissions can also be calculated based on the amount of petrol used. Emissions
should be less than 80 g per kilometre by 2020 if we want to achieve the targets for
saving the climate.
Not only the type of car but also the way in which it is driven has a major impact
on how much petrol is consumed. The following energy-saving tips will enable
savings of up to 30%:
■ Check tyre pressure frequently and pump up tyres at a minimum to the pres-
sure for a fully loaded vehicle recommended by the car’s manufacturer.
■ Change gear as early as possible, drive at a uniform speed and always antici-
pate what is ahead of you.
■ Restrict maximum speed on motorways.
■ Do not transport unnecessary ballast or unneeded roof luggage racks.
■ Switch off air conditioning and other energy-consuming devices when they
are not needed.
■ For short distances try to walk or go by bicycle.

Determining Carbon Dioxide Emissions Based on Petrol Usage

Litre petrol g CO2 g CO2


Petrol x 23.7 =
100 km Litre petrol/100 km

Litre diesel g CO2 g CO2


Diesel x 26.5 =
100 km Litre diesel/100 km

The capacity utilization of a car has the biggest effect on energy use. On average
three to four seats in a car are empty when it is being driven. With four people in
a vehicle and careful driving, the carbon dioxide emitted per person can be even
lower than with a train.
Capacity utilization also plays an important role with other forms of transport. As
the capacity utilization of public local transport is on average lower than that of
long-distance travel, the rate of carbon dioxide emissions is also higher. A coach
filled to capacity is particularly economical in energy use. In contrast, heavy fast
trains such as France’s TGV can consume surprisingly high amounts of energy.
Air traffic produces the highest carbon dioxide emissions. Long-haul flights cause
the most harm to the climate, because the exhaust gas of planes is more damaging
when emitted at high altitudes. If this factor is calculated into carbon dioxide emis-
sions, these rise to 400 g per kilometre.
The car industry is investigating numerous solutions to enable climate-neutral
mobility in the medium and long term. These include:

60
3.3 Industry and Co – Everyone Else is to Blame

■ Biofuels
■ Electric cars using renewable electricity
■ Transmissions based on renewable hydrogen
Although the natural gas vehicles currently being marketed emit slightly lower
levels of carbon dioxide than before, they really do not offer an alternative for
protecting the climate worldwide. Biofuels may be able to replace oil relatively
easily and are sometimes added to conventional fuels. However, the problem is that
not enough biomass is available to provide an adequate supply. Even if all the
farmland in places like Germany and Britain were used to plant the raw materials
needed for biofuels, this would not come close to covering the current fuel needs
of motor vehicles in those countries. On closer inspection, even renewable hydrogen
poses some problems. (Biofuels and hydrogen will be covered in detail in Chapters
12 and 13, respectively.)
Until now the main obstacle to electric cars has been limited battery life. Long
charging times and the short distances that can be driven on a single charge are
preventing the widespread use of these cars. Considerable progress is currently
being made in battery development, which could eventually make electric cars a
viable option. However, these cars cannot offer a real alternative for saving the
climate unless the electricity used to charge the batteries is also sourced from renew-
able power plants.

3.3 Industry and Co – Everyone Else is to Blame

It’s widely believed that industry and the energy companies are mostly to blame for
greenhouse gas emissions and so private individuals can’t do much about the
problem. But looking at the situation more closely, one sees that they are only
complying with what the customers want. Therefore, in the final analysis it is the
consumers who are responsible for the emissions because of the products they are
demanding.
If all electricity customers were to change to suppliers or tariffs that only offer
electricity from renewable energies, the power supply would very quickly be free
of carbon dioxide. The only problem the energy suppliers would then have is being
able to build enough renewable power plants as quickly as possible to accommodate
all the sudden demand.
Due to their choice of products, consumers are failing to put pressure on the right
places for a sustainable type of economy. The production of any type of product
– whether a food item or consumer goods – requires energy and therefore causes
carbon dioxide emissions. The higher one’s personal consumption, the higher the
use of energy and the higher the carbon dioxide emissions. Yet even those who
swear off consumption completely will not be able to reduce their energy needs to
zero, because food production produces quite high emissions.

61
3 From Wasting Energy to Saving Energy and Reducing Carbon Dioxide

Each consumer can have an important impact on indirect energy consumption


through product selection:
■ Try to buy only high-quality and durable products.
■ Give preference to regional products.
■ Select products that require less energy during the manufacturing process
and produce lower emissions.
■ Give preference to companies with environmentally friendly policies.

3.4 The Personal Carbon Dioxide Record

Even if users follow all the energy-saving tips provided, they will probably find that
the carbon dioxide emissions they are responsible for are still quite high. This
section enables readers to do a self-assessment by explaining how to calculate the
carbon dioxide emissions an individual can personally cause.

3.4.1 Emissions Caused Directly by One’s Own Activities


The easiest way to determine which emissions are the result of one’s own activities
is to look at how much oil, natural gas, petrol and electricity one has used. It is rela-
tively simple to establish the amount of energy consumed per year by checking bills
and invoices.
■ Annual electricity bill
■ Heating bills
■ Number of kilometres driven and average petrol consumption
■ Kilometres travelled using public transport
■ Number of miles flown
Based on this information, the following calculation method can be used to deter-
mine the emissions resulting from one’s own activities.
Specific consumption values may have to be adapted to a person’s own particular
circumstances. In electricity generation in Germany on average 0.616 kg of carbon
dioxide are produced per kilowatt hour (kg CO2/kWh) of electric power. The value
in the USA is roughly the same. Those who get green electricity from renewable
power plants can cut their emissions from electricity consumption to zero.
Emissions from heating can vary considerably. If heat is generated electrically, the
emissions are the same as for the electricity. With modern natural gas heating
0.2 kg CO2/kWh of heat is created; with modern oil heating it is 0.28 kg CO2/kWh.
With old, inefficient heating systems the emissions can rise from 0.25 to 0.35 kg CO2/
kWh. Heating with biomass at most generates indirect carbon dioxide emissions due
to processing and transport. With wooden pellets the value is around 0.06 kg CO2/kWh.
The carbon dioxide emissions specific to a person’s own car can be calculated on
the basis of average fuel consumption (see Planning Help, above). With air travel

62
3.4 The Personal Carbon Dioxide Record

an emissions calculator makes it relatively easy to calculate emissions precisely,


according to air route.

A. Calculation of Annual Private Direct Carbon Dioxide Emissions

■ https://www.atmosfair.com/index.php? Atmosfair emissions calculator


id=5&no_cache=1&L=3 for air trips

3.4.2 Indirectly Caused Emissions


In addition to the emissions a person has caused directly, everyone is also indirectly
responsible for other emissions. Energy is consumed in the production, processing
and transport of foodstuffs, consumer goods and other products, and this in turn
causes carbon dioxide emissions.
The public sector, including government offices, schools, the police, the fire brigade
as well as the departments responsible for roadwork, also needs energy and therefore
causes carbon dioxide emissions. About one-sixth of emissions are attributed to
public consumption. This is an area where there are not many possibilities to make
personal reductions.
About 1.2 tons of carbon dioxide per person are produced every year due to our
individual food consumption. Figure 3.10 compares the greenhouse gas emissions
of different foodstuffs. In addition to carbon dioxide, the chart takes into account

63
3 From Wasting Energy to Saving Energy and Reducing Carbon Dioxide

butter
beef
cheese
frozen french fries
poultry
pork
eggs
curd cheese
yoghurt
milk
pasta
brown bread
conventional f arming
vegetables, tinned
organic f arming
vegetables, frozen
potatoes, fresh
vegetables, fresh

0 2 4 6 8 10 12 14 16 18 20 22 24
kg greenhouse gas emissions per kg food product

Figure 3.10 Greenhouse gas emissions converted into carbon dioxide equivalents for
the production of different foodstuffs. Data: (Fritsche and Eberle, 2007).

other greenhouse gases such as methane and nitrous oxide and converts these values
into carbon dioxide equivalents according to how harmful they are to the environ-
ment. Beef, butter and cheese do not come out favourably. Ruminant cows release
large amounts of methane that have a high greenhouse effect. Subsequent freezing
of the meat also adds to the energy toll, so that people who prefer frozen meat to
fresh meat increase greenhouse gas emissions by another 10 to 30%.
The amount of processing that food requires also has a considerable effect on the
climate. For example, fresh potatoes hardly tip the scales of greenhouse gas emis-
sion. However, the energy-intensive further processing required to make frozen
French fries increases the greenhouse gas emissions contained per kilogram of food
almost 30 times.
As women normally eat less than men and older people eat less than young people,
clear differences exist from person to person. Going on a permanent diet is certainly
not the right approach to reducing carbon dioxide emissions. However, people who
buy seasonal products from local organic farms, reduce their consumption of meat
and largely avoid tinned and frozen foods can significantly reduce emissions resulting
from food intake. Reductions of up to 30% can easily be achieved this way. However,
those who need a large number of calories because they do a lot of physical work or
sport, and then obtain these calories mainly from frozen products and fast foods,
could find they double their greenhouse gas emissions just from what they eat.
Personal consumption is responsible for around 2 to 3 tons of carbon dioxide. This
is mostly from the production, storage and transport of all products that do not fall
under the category of food supply. Included in this group are clothing, furniture,
machines and equipment, paper products, cars and housing.

64
3.4 The Personal Carbon Dioxide Record

Each German uses about 240 kg of paper each year, each Briton 200 kg or more,
and each US citizen around 300 kg. The production of 1 kg of new paper produces
about 1.06 kg of carbon dioxide. This means that almost 10% of personal consump-
tion is attributed to the use of paper. The energy consumption with the production
of recycled paper drops by around 60% and the carbon dioxide emissions by about
16%. Whereas recycled paper was very popular in the 1980s, many distributors no
longer carry it at all – even though its benefits to the environment have not changed.
The reason is weak demand on the part of the consumer because of unfounded
claims that recycled paper destroys printers and copiers and is even damaging to
one’s health.
The rule generally is that anyone who is a consumer is also contributing to high
emissions. On average around 4 to 5 kg of carbon dioxide are created for each euro
spent by consumers. Here too there are many ways to make reductions. High-quality
long-life products almost always produce low amounts of carbon dioxide. Although
they are usually more expensive initially, in the long run they actually protect the
wallet as well as the environment because of their long useful life. Increasing the
use of natural materials may not always be less expensive but it does usually reduce
greenhouse gas emissions. For example, building a new solid house is almost as
expensive as putting up a prefabricated house. However, if wood frame construction
is used for the prefabricated house, the additional use of wood as a building material
will save around 20 to 30% in carbon dioxide emissions. As a result, the savings
potential quickly runs into several tons.
In the following calculation method, the indirectly caused greenhouse gas emissions
can be inserted in the spaces shown.

B. Calculation of Annual Indirect Carbon Dioxide Emissions

3.4.3 Total Emissions


Once the personal direct and indirect emissions have been determined, they can be
added up to get a figure for all personal carbon dioxide emissions. Somewhat more
than 10 tons of carbon dioxide are generated on average per head in Germany each

65
3 From Wasting Energy to Saving Energy and Reducing Carbon Dioxide

year. In Great Britain the figure is 8.9 tons and in the USA more than 19 tons. If
other greenhouse gases, such as methane and nitrous oxide, are taken into account
and converted into carbon dioxide equivalents based on how damaging they are to
the climate, then the total greenhouse gas emissions in Germany increase to more
than 12 tons per head per year. These results are, however, limited to carbon dioxide.

C. Calculation of Annual Total Personal Carbon Dioxide Emissions

An even more precise calculation of personal emissions can be made using the
emission calculators on the Internet. Figure 3.11 enables a comparison of one’s own
carbon dioxide emissions with the reduction targets and emissions of other coun-
tries. The colour scale in the figure shows how the emissions should be evaluated.

Figure 3.11 Scale of emissions of carbon dioxide per head and year.

If possible, global carbon dioxide emissions should be halved by the middle of the
century. This means that in the meantime the activities of each inhabitant are
allowed on average to produce less than 2000 kg of carbon dioxide.

■ https://www.lfu.bayern.de/luft/fachinformationen/ Online CO2 calculator for


co2_rechner Germany
■ https://www.epa.gov/climatechange/emissions/ US Personal Emissions
ind_calculator.html Calculator
■ https://www.nef.org.uk/greencompany/ UK Simple Carbon
co2calculator.htm Calculator

66
3.5 The Sale of Ecological Indulgences

3.5 The Sale of Ecological Indulgences

We may be able to make major reductions in our own direct and indirect emissions.
However, in most industrialized countries it will still be relatively difficult to move
emission levels into the ‘green’ zone. Some emissions, such as those related to
public consumption, lie well beyond one’s personal sphere of influence. Other
reductions can only be achieved through radical changes in lifestyle or relatively
high investment.
People who still want to make further reductions can use the emissions trade to
compensate for their own emissions. This idea is even being practised on a large
scale by states that want to reduce their overly high emissions as part of their com-
mitment at the international level. The modern emissions trade is slightly reminis-
cent of the Medieval church’s practice of selling indulgences. However, as the
expression ‘sale of indulgences’ has negative connotations, the term joint imple-
mentation (JI) was invented.
Whatever is planned on a grand scale at the state level, and has partially already
been implemented, can also be done in the private sphere. For instance, you could
give your neighbour an inexpensive low-energy light bulb as a gift. This bulb could
save up to 300 kg of carbon dioxide emissions over its lifetime. If you gave away
enough low-energy light bulbs, you could conceivably save your total emissions in
another area – at least theoretically. The actual personal emissions continue to
accumulate and would then have to be transferred to the emission results of your
neighbour. In pure calculation terms, the neighbour’s emission results would remain
constant. However, if the neighbour’s emissions are also too high, he will not
have an easy option for reducing his emissions himself. Furthermore, due to the
savings in electricity costs, the neighbour will end up with more money in his
household budget. If he then invests this money in a couple of litres of petrol for
an extra jaunt in the car, then the whole exercise has ended up being highly
counterproductive.
The situation is different if a low-energy light bulb is given to a school in a devel-
oping country with emissions in the green zone. In this case the gift of the light
bulb will reduce already low emissions even further. Naturally the prerequisite even
here is that the school is not in a position to buy the bulb itself. If the school is
already planning to buy a low-energy bulb as an example of reducing electricity
costs, the bulb given as a gift will not really represent a savings effect. Practised on
a large scale, this type of investment is called Clean Development Mechanism
(CDM).
The following criteria should be included to ensure the success of climate protection
projects:
■ Carbon dioxide reductions through financing renewable energy plants or
energy-saving activities.
■ Implementation of an additional measure that would not necessarily have
been carried out during the lifetime of a project.

67
3 From Wasting Energy to Saving Energy and Reducing Carbon Dioxide

■ Guarantee of successful project implementation and plant operation during


the entire project.
■ Ongoing development and technology transfer after project completion.
It is difficult for a private individual to meet these criteria. However, various com-
panies offer professional services for implementing climate-protecting projects.
Even major projects are possible if a large number of customers join forces. However,
an independent auditing authority should always be available to monitor that the
process is being followed correctly (Figure 3.12). Atmosfair is an example of a
company that offers a programme for offsetting emissions from plane trips. One of
the projects in their programme involved replacing diesel stoves in large kitchens
in India with solar mirrors.

Figure 3.12 Principle of private trade in emissions.

Currently about 20 to 30 euros should be allocated to compensate for the emission


of 1000 kg of carbon dioxide with clean development projects. Thus it will cost a
person 10 euros each month to prevent the emission of 5000 kg per year. Protecting
the climate is really not that expensive.

■ www.atmosfair.com Different suppliers for reducing carbon dioxide emissions


■ www.co2ol.de through climate-protection projects
■ www.carbonneutral.com

Naturally emissions can be offset through renewable energy projects at home. In


countries such as Germany, for example, anyone who invests in a wind farm in a
good location or installs a large photovoltaic system on an optimally orientated roof
can even make a small profit. Germany’s renewable energy law makes this possible.
This law establishes the subsidy levels for electricity from renewable energy sup-
pliers, such as wind farms, photovoltaic systems and biomass, geothermal and
hydropower plants. These subsidy levels are normally higher than regular subsidies
for electricity from conventional power plants. The energy supply company into
whose grid the plants feed their electricity is allowed to split the extra costs among
all its electricity customers (Figure 3.13). This means that all customers – whether

68
3.5 The Sale of Ecological Indulgences

Figure 3.13 Principle of financing renewable power plants through the renewable energy law
in Germany.

they want to or not – are helping to finance the renewable energy plants. The burden
on the individual customer is therefore kept to a minimum. As a result of the renew-
able energy law in Germany, in 2008 the allocated extra costs of all renewable power
plants amounted to around 1.10 euros per month per person.
The German renewable energy law is the most successful instrument worldwide for
promoting the rapid development of renewable energy. It aims to involve everyone
equally in the social task of restructuring the energy sector – be it as investor, opera-
tor of renewable energy plants or financier for the reimbursement of electricity from
existing plants.

69
4
4 Carbon-Free Energy – Vision
or Utopia?
Most of us who have calculated our personal carbon footprint have made the sober-
ing discovery that our personal emissions are far too high. As the previous chapter
shows, savings can be made in different areas. However, in most cases the pos-
sibilities open to individuals for reducing greenhouse gas emissions still do not make
them nearly green enough. Investments in renewable energies can help to reduce
emissions somewhere else. The climate can only be saved if in the long term all
countries on earth reduce their greenhouse gas emissions to almost zero. However,
there are many among us who just cannot imagine life without oil, natural gas
and coal.
Yet a mere 300 years ago renewable energies made up the earth’s entire energy
supply. It is quite certain that the world’s energy supply will once again be com-
pletely carbon-free 200 years from now. By then the last deposits of fossil energy
sources will have been exhausted. As a result of an almost 500-year history of fossil
energy use, the climate would have totally collapsed by then. If we want to prevent
this happening, we must convert to carbon-free energy supplies long before then.
At most we have about 100 years in which to do this (Figure 4.1).

4.1 Options for Carbon-Free Energy Supply

4.1.1 Efficient Power Plants – More Power with Less


Carbon Dioxide
Representatives from the electricity sector like to quote efficiency gains in power
plants as important milestones in the effort to achieve efficient climate protection.

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

70
4.1 Options for Carbon-Free Energy Supply

100%

90%
nuclear power
80% development 1:
primary energy demand climate protection
70%
share of world

fossil development 2:
60% fuels depletion of
all reserves
50%

40%
traditional modern
30% renewable renewable
energy suppy energy supply
20%

10%
0%
1000

1100
1200

1300
1400

1500
1600

1700
1800

1900
2000

2100
2200

2300
2400

2500
2600

2700

2800
2900

3000
year

Figure 4.1 Development of the share of different energy sources in world primary energy
demand and two future development possibilities.

In reality, enormous quantities of carbon dioxide could be saved if old ailing and
poorly performing power plants were replaced with ultra-modern ones. Whereas old
brown coal plants usually reached an efficiency of less than 40% in electricity pro-
duction, new plants can easily reach 43%. Technically, efficiency of 48 to 50% is
possible.
For example, with an efficiency of around 35%, the Jänschwalde power plant near
Cottbus in Germany is one the oldest and most inefficient large power plants in
eastern Germany (Figure 4.2). With an output of about 25 tons of carbon dioxide
per year, this plant alone produces about 3% of all carbon dioxide emissions caused
by energy use in Germany.

Figure 4.2 Jänschwalde brown coal power plant and opencast mining near Cottbus.

71
4 Carbon-Free Energy – Vision or Utopia?

The newly built BoA 2/3 Neurath power plant on the edge of the city of Grevenbroich
in Germany uses optimized technology and is an example of the efficiency that can
be reached in a power plant. The 2.2 billion euro plant with efficiency of over 43%
will be replacing some older power plants in 2010, thereby saving around 6 million
tons of carbon dioxide per year. The main problem, however, is that the power plant
will continue to emit around 14 million tons of carbon dioxide every year. As an
example, this is far higher than the total carbon dioxide emissions caused by the 35
million people in Kenya. With an average lifetime of 40 years, even a large, efficient
power plant quickly develops into an obstacle to progress when it comes to effective
climate protection. By the time the Neurath power plant is supposed to be discon-
nected from the grid in 2050, industrialized countries like Germany should have
achieved an 80% savings in carbon dioxide emissions overall. This is a level that
not even the most efficient power plants can achieve. As with nuclear power plants,
it is even possible that fossil power plants will be taken out of service earlier than
planned to enable targets for climate protection to be met.
Another problem with coal-fired power plants is that they are relatively difficult to
regulate. Brown coal-fired plants are usually designed for a base load. This means
they work optimally when they are providing constant output for long periods of
time. However, due to the increased use of wind power and solar energy plants,
fluctuations in power are increasing in the grid. Building new coal plants does not
make a great deal of sense for this reason alone.

4.1.2 Carbon Dioxide Sequestering – Away with


Carbon Dioxide
As even a medium-term increase in efficiency in fossil power plants does not offer
an alternative to the supply of climate-compatible energy, supporters of fossil energy
plants have looked for arguments and ways to avoid coming under even more fire
because of the damage caused to the climate.
Carbon dioxide sequestration is suggested as a way out of the dilemma. The idea
behind it is simple and at first glance also appears highly plausible. The idea is that
power plants of the future will no longer emit the carbon dioxide that occurs with
the burning of coal and natural gas into the atmosphere but instead capture and store
it in a safe place (Figure 4.3).
The following options are promising for the safe disposal of carbon dioxide:
■ Manufacture of carbon-based building materials by the construction material
industry.
■ End-storage underground in depleted fossil deposits, in aquifers and in
salt beds.
■ Compression or dissolution in the ocean.
■ Bonding through special algae in the ocean.
However, a critical examination of the options available produces some doubts that
put a question mark on the disposal of carbon dioxide in general. The technology
for sequestration and storage is still at the research stage. It will be several years

72
4.1 Options for Carbon-Free Energy Supply

Figure 4.3 Options for end-storage of sequestered carbon dioxide.

until it is actually available to commercial installations. Currently only studies and


prototype facilities exist. Commercial introduction is envisaged for the year 2020
(Vattenfall Europe, 2006). However, those promoting climate protection are
demanding carbon dioxide savings of 40 to 50% by that time. Just from the stand-
point of time this will not be possible with carbon dioxide sequestration. Furthermore
it will be very difficult – if even possible – to upgrade the old power plants that
would have been built in the meantime.

Carbon Dioxide-Free Coal-Fired Power Plants – Not Really Free


of Carbon Dioxide
With the development of technologies for the sequestration and storage of
carbon dioxide, grandiose claims are made that power plants will one day be
carbon dioxide-free. Strictly speaking, however, even carbon dioxide retention tech-
niques can never make a fossil power plant free of carbon dioxide. The burning of fossil
energy sources in these power plants inevitably produces carbon dioxide. So-called
carbon dioxide-free fossil power plants only separate carbon dioxide from other combus-
tion gases so that it can be stored ultimately in a concentrated form outside the atmos-
phere. Carbon dioxide also escapes inadvertently during fuel processing, sequestration
and storage. In the long term up to 10% of carbon dioxide, and in individual cases even
more, reaches the atmosphere again.
The term ‘carbon dioxide-free coal-fired power plant’ is therefore misleading. A company
in the solar sector actually sued an energy supply company for misrepresentation over
its claims that carbon dioxide-free power plants were being built in Germany. The court
ruled in favour of the solar company. Whether or not fossil power plants with carbon
dioxide separation are carbon dioxide-free in legal terms is likely to be examined by
other authorities.

73
4 Carbon-Free Energy – Vision or Utopia?

Controversy also surrounds end-storage in the ocean. The possibility exists that the
carbon dioxide will escape again into the atmosphere after a short period or will
have an extreme but not yet identifiable effect on the ecological system of the
oceans. This is an area that generally still requires extensive research. Other types
of storage areas are not available everywhere or would quickly be exhausted. If all
power plants in the world were required to dispose of their carbon dioxide in a way
that would not damage the climate, it would take an enormous logistical effort to
transport the carbon dioxide over what could sometimes be thousands of kilometres
to end-storage depositories.
Probably the most important argument against global sequestration of carbon
dioxide is the economic viability of doing so. Carbon dioxide sequestration gener-
ally reduces the efficiency of a power plant. Added to this is the cost of the transport
and end storage of the carbon dioxide. It is difficult at present to estimate the exact
cost involved. Many estimates predict that the cost increases for electricity from
fossil power plants would be up to double present values (Intergovernmental Panel
on Climate Change, 2005). For many countries in the world this would rule out even
the possibility of carbon dioxide sequestration. In contrast, cost estimates for renew-
able power plants show that in many locations they would prove to be more eco-
nomical than fossil power plants within a relatively short period – regardless of
whether or not they have carbon dioxide sequestration.

■ www.ipcc.ch/ipccreports/srccs.htm Background information on CO2


separation and storage

4.1.3 Nuclear Energy – Squeaky Clean


Almost every discussion on climate ends up looking at nuclear energy as a possible
saviour. The arguments against nuclear energy have already been explained in detail
in previous chapters of this book. Even if the very conflicting evaluated risks of
nuclear energy are disregarded, this solution does not offer an option for effective
climate protection in the medium term. The main reasons for this are:
■ The proportion of nuclear energy in the global primary energy supply is only
around 6%. In terms of final energy consumption, the proportion is consider-
ably lower.
■ The amount of uranium that can be extracted economically is very limited.
The price of uranium has already increased significantly in recent years.
■ Nuclear energy plants are difficult to regulate and therefore are not suitable
for joint operation with wind power and solar plants.
■ Although nuclear energy is practically carbon-free it gives rise to other risks,
such as nuclear accidents, terrorist attacks and the unresolved question of
end storage.
■ Nuclear fusion will not be ready for use for several decades and thus too late
to save the climate; it is also difficult to regulate and is extremely expensive.

74
4.1 Options for Carbon-Free Energy Supply

4.1.4 Combined Heat and Power – Using Fuel Twice


When it comes to increasing efficiency, combined heat and power (CHP) is often
promoted as a promising candidate. Using electricity generated by conventional
steam turbine power plants that burn brown and hard coal, the level of efficiency is
between 35 and 45%. Modern gas and steam turbine power plants that use natural
gas achieve up to 60%. However, this means that at least 40% of the primary energy
fizzles out unused through the cooling tower of the power plant.
CHP, or cogeneration, plants use the heat from electricity generation effectively and
are able to exploit up to 90% of the fuel. As a result, in an optimal case less carbon
dioxide is produced than when electricity and heat are generated separately.
Many comparisons of cogeneration power plants that generate heat and electricity
separately using fossil fuels show possible carbon dioxide savings of up to 50%.
However, these comparisons usually pit CHP plants against antiquated electricity
plants. If the comparison is made on the basis of optimal functioning plants on both
sides, the carbon dioxide savings are reduced to a meagre 15 to 20% (Figure 4.4)
– too little to save the climate. Furthermore, these savings are only possible with
optimal plant operation. For example, in summer when a cogeneration plant is only
supposed to generate electricity but not heat, it will have great difficulty in even
coming close to the dream efficiency level of 90%. A cogeneration plant can some-
times end up producing even more carbon dioxide than a straightforward electricity
power plant.

Figure 4.4 Comparison of primary energy requirements and carbon dioxide emissions
between power-heat coupling and separately generated heat and electricity in modern power
plants.

If, on the other hand, a sufficient requirement for heat exists over the entire year,
cogeneration plants can help to reduce carbon dioxide. With cogeneration plants
that use fossil fuels the savings for effective climate protection are too low. Those
that use renewable energy sources, such as biomass and geothermal energy, are
totally carbon-free and can continue to accelerate the switch to carbon-free energy
supply.

75
4 Carbon-Free Energy – Vision or Utopia?

4.1.5 Saving Energy – Achieving More with Less


As the last chapter showed, the options for saving energy are enormous – at least
in industrialized countries like Germany. The situation is different in developing
countries. Someone who does not even own an electric light bulb and lives in a
house without heating will be in no position to save energy through energy-saving
bulbs or building insulation. Figure 4.5 shows that per capita energy requirements
generally increase with the per capita gross domestic product (GDP), which roughly
mirrors the prosperity of a country.

500
Iceland
450 Bahrain
Kuwait
400 Arab
primary energy demand

Trinidad/Tobago Emirates Canada


350 USA
in GJ per capita

Brunei NL Antilles Finland


300
Saudi Arabia Singapore Sweden Norway
China
250
Belgium Australia TREND
India Oman Netherlands
200
Russia KoreaTaiwan F
Kazakhstan Estonia CZ NZ D GB Austria
JP Ireland
150 SL CY E DK
South Af rica SK Switzerland
P
100 H GR IS Italy Hong Kong
Malta
50 Argentina
Uruguay
0
0 5 10 15 20 25 30 35 40
GDP (PPP) in 1000 US$2000 per capita

Figure 4.5 Per capita primary energy consumption based on gross domestic product (GDP)
according to the purchasing power parity method (PPP) for different countries.

The differences between countries with the same GDP per inhabitant are consider-
able. Whereas the GDP per inhabitant in Canada and in Switzerland is almost
identical, a Canadian consumes far more than double the primary energy of a Swiss
person. In addition to the way people live and how they handle energy, the climate
conditions and industrial structure of a country also play an important role in energy
consumption. However, having high energy needs does not automatically equate to
high greenhouse gas emissions. Although the per capita energy requirement of
Iceland is clearly higher than that of Canada, its carbon dioxide emissions are sub-
stantially lower. This is due to the fact that Iceland covers most of its energy require-
ments carbon-free using hydropower and the natural heat of the earth.
Many countries in the world have a low level of prosperity and, consequently, a
low demand for energy. Once the per capita energy requirements of China and India
reach the same levels as Great Britain or Germany, this alone will more than double
global demand. Considering the high growth rates of both countries, this would

76
4.2 Renewable Energy Sources – No End to What is Available

seem to be only a matter of time. If the industrialized countries halve their energy
needs in the meantime and the developing countries follow our example, this would
help considerably in slowing down the rise in energy requirements.
Along with the growing per capita energy requirements in developing and emerging
countries, the continuous rise in the world’s population is also adding to the steady
increase in energy demand. Between 1960 and 2000 the world population doubled,
and energy requirements tripled. Worldwide energy demand has been increasing
almost as quickly as the population since 1980 (Figure 4.6). If the world population
continues to climb at this rate and reaches nine billion by 2050, this alone will mean
a 50% increase in energy demand – even without any increase in per capita energy
requirements.

7 500

_
world primary energy demand in EJ
450
world population in billion people

6
400
5 350

300
4
250
3
200

2 150

100
1
50

0 0
1950 1960 1970 1980 1990 2000 year 2010

Figure 4.6 Development of worldwide primary energy demand and increase in world
population.

These facts clearly show that energy-saving measures are enormously important in
at least putting a brake on the increasing demand for energy worldwide. However,
these measures alone will not be enough to achieve a major reduction in worldwide
greenhouse gas emissions over the next 50 years. Along with implementing every
conceivable measure to save energy, the most important thing will be to ensure that
those energy requirements that cannot be eliminated are at least carbon-free. There
is also a comprehensive solution to this: renewable energies.

4.2 Renewable Energy Sources – No End to


What is Available

The options for supplying climate-compatible energy discussed above offer only
limited possibilities for reducing carbon dioxide. The situation is totally different
with renewable energy sources: these offer us almost unlimited potential.

77
4 Carbon-Free Energy – Vision or Utopia?

Each year the sun radiates 1.5 quintillion kilowatt hours of energy towards the earth.
The atmosphere swallows up around 30% of this energy but over one quintillion
kilowatt hours are still able to reach the earth’s surface. Our current primary energy
needs are around 125 trillion kilowatt hours worldwide. By the way, a quintillion
is a 1 with 18 zeros, and a trillion has 12 zeros. Therefore, the amount of energy
that reaches the earth’s surface from the sun each year is 8000 times more than the
total primary energy requirement of the world. So we only need to use about one
hour’s worth of the solar energy that reaches the earth’s surface in order to cover
the energy needs of the whole of mankind for a whole year.
Natural occurrences convert some of the sun’s energy into other renewable forms
of energy, such as wind, biomass and hydropower. In addition to these energy forms,
we are also able to use the natural heat of the earth as well as tidal power derived
from the motion of the moon in conjunction with other planets. All sources of
renewable energy combined exceed the total fossil and nuclear fuels available on
earth many times over (Figure 4.7). In less than one day the sun radiates more energy
to the earth’s surface than we could ever use if we were to burn all the oil reserves
available on earth.

Figure 4.7 Comparison of annual renewable energy available and global primary energy
requirement with the total existing conventional energy sources on earth.

In discussions on climate, some critics question whether renewable energies are


even able to cover our energy requirements. But a brief glance at the facts just
mentioned shows that these doubts can safely be cast aside. The variety of possible
uses of renewable energies is enormous. A number of different types of power plants
can provide almost any amount of electricity, heating or fuel desired (Figure 4.8).
The following chapters of this book will present in detail the key technologies for
using renewable energies.

78
4.3 Options for Protecting the Climate

Figure 4.8 Sources of and possibilities for using renewable energies.

4.3 Options for Protecting the Climate

A typical scenario for protecting the climate of a country like Germany should first
show the options available to make the energy supply carbon-free quickly, and
identify the role renewable energies can play in this process. Compared to other
countries, the possibilities for using renewable energies in Germany are far from
optimal. However, if a highly populated industrialized country like Germany with
only moderate potential could succeed in using renewable energies to supply its
needs, then it should not be a problem at all for other countries. Germany is a trail-
blazer here, which offers many advantages in the long term. The country is already
reaping the benefits of its role, as renewable technologies are developing into a big
export success for German industry.

4.3.1 Down with Primary Energy Needs


The total primary energy use in Germany has to fall for climate protection to be
effective. Primary energy includes original forms of energy, such as coal, oil, natural
gas and uranium – most of which are used in the energy sector. Some is also used
for non-energy related purposes in other areas. For example, large quantities of oil
end up in the plastics industry where it is used to make garden chairs, ballpoint pens
and nylon tights. The majority of these products ultimately land in a rubbish tip.
The rubbish that is burnt in refuse incinerators then produces carbon dioxide, which
in turn contributes to the greenhouse effect. Although alternative raw materials
such as natural bioplastics are now available, it is expected that during the next
few decades fossil raw materials will play a major role in non-energy use. To be

79
4 Carbon-Free Energy – Vision or Utopia?

effective in protecting the climate, Germany should be reducing its greenhouse gas
emissions by 80% by 2050. However, if the non-energy part is largely left unchanged,
this means that the energy sector will have to convert almost completely to renew-
able energies by then.
In energetic primary energy consumption in 1990 the combined fossil and nuclear
energy share was almost 99%. In the same year renewable energies only contributed
a maximum of 1% to energy requirements. The share covered by renewable energies
then rose to over 7% by 2008 (Figure 4.9). So it is still a long road to 100% renew-
able energy supply.

16000
non-energetic
14000 f ossil f uels
primary energy demand in PJ

nuclear power
12000 hydrogen
heat pumps
10000
geothermal
biomass
8000
solar import

6000 solar thermal


photovoltaics
4000 wind power
hydropower
2000

0
1990

1995

2000

2005

2010

2015

2020

2025

2030

2035

2040

2045

2050

Figure 4.9 Possible development of renewable energy share in total primary energy
consumption in Germany.

Between 1990 and 2008 nuclear energy had a share of between 10 and 14%. As a
result of its opt-out decision, Germany is expected to end its use of nuclear energy
between 2020 and 2025. Until then renewable energies will easily be able to replace
the dwindling capacities. Whereas the primary energy requirement remained largely
constant between 1990 and 2005, it will fall steadily because of more efficient
energy use and a drop in population numbers by 2050. As power plant efficiency
is not included in the calculation of primary energy consumption for electricity from
wind, hydropower and photovoltaic plants in contrast to coal-fired and nuclear
plants, primary energy consumption is also decreasing due to statistical effects.
Overall the primary energy requirement could be more than halved in Germany.
The changes taking place in the energy sector in Germany are about 10 to 15 years
too late to meet the targets for climate protection in 2020. Instead of the recom-
mended 50% reduction in greenhouse gas emissions, a figure closer to 30% is more
realistic. Keeping the country’s nuclear power plants open for longer would also
not help to meet the targets set for 2020. Instead it would delay a reorganization of
the energy sector without making a necessary contribution towards climate protec-

80
4.3 Options for Protecting the Climate

tion. However, the targets for 2050 can easily be achieved if a consistent course is
followed until then.
Germany’s energy in 2050 will probably come from a broad renewable energy mix
(Figure 4.10). Because of the country’s geographical situation, the potential for
hydropower will be quite low. In addition to indigenous renewable energies, the
import of reasonably priced solar power from sunny regions will probably also
contribute towards providing safe and cost-effective energy.

non-energetic
demand hydro power
16 % 1%
wind power
renewable 9%
hydrogen
14 % photovoltaics
7%

heat pumps
8% solar thermal
10 %

solar import
geothermal
7%
12 %

biomass 16 %

Figure 4.10 Possible share of renewable energies in primary energy consumption in


2050 in Germany.

4.3.2 Electricity Generation Totally Without Nuclear and


Fossil Power Plants
Whereas a big decline is expected in overall primary energy use in Germany over
the next few decades, electricity consumption will probably remain constant. The
potential to reduce electricity requirements is also high. However, the trend towards
always having more and more new electrical appliances and the increased use of
heating pumps and electric cars will cancel out this potential.
The changeover to renewable energy supply has already begun in the electricity
sector. In 1990 hydropower, with its share of around 3%, was the only energy
source worth mentioning in this area. In the next 15 years this was followed by
wind power, biomass use and, more recently, photovoltaics. These had already
amassed a 15% share by 2008. Renewable energies should be able to replace nuclear
energy by 2020, and by 2050 virtually cover all of Germany’s electricity needs
(Figure 4.11).
There may be little potential for hydropower in Germany, but wind power and
photovoltaics do offer good possibilities. Theoretically, either wind power or pho-
tovoltaics alone could potentially satisfy the country’s entire demand for electricity.

81
4 Carbon-Free Energy – Vision or Utopia?

700

600
gross electricity demand in TWh

500 f ossil f uels


nuclear power
import (renewable)
400
geothermal
biomass
300
photovoltaics
wind power
200 hydropower

100

0
1990

1995

2000

2005

2010

2015

2020

2025

2030

2035

2040

2045

2050
Figure 4.11 Possible development of share of renewable energies in gross power
consumption in Germany.

In practice, however, this does not make much sense, because large and expensive
storage would be necessary due to the fluctuating availability of wind power and
solar radiation over the course of the year.
If different renewable energies are combined in a sensible way, then only small
storage systems are necessary. At the same time wind power and photovoltaics could
be providing around half the electricity supply by 2050. The development of biomass
power plants is also limited, but biomass will clearly exceed the share now provided
by hydropower. In the medium term geothermal power plants will also be able to
make a contribution. Supplemented with the import of reasonably priced electricity
from renewable power plants, renewable energies will enable demand to be covered
with carbon-free electricity by 2050. Chapter 7 is therefore devoted to solar-thermal
power plants, which offer an important option to the import of electricity. Fossil
power plants and nuclear power plants will no longer be needed to provide safe and
climate-compatible electricity supply in 2050.

4.3.3 Insulation and Renewable Energies to Provide Heat


It may be difficult to cut down on electricity consumption, but substantial reductions
are possible when it comes to heating. The energy-saving possibilities offered by
the buildings described in Chapter 3 make it probable that heating needs will be
halved by 2050 compared to 1990 (Figure 4.12). What is important is that govern-
ment strictly enforces modernization measures in existing buildings.
In 2007 the share of renewable energies in heat supply was around 6%. Among
these renewable energies, biomass – i.e. firewood – had the greatest share by a long
shot. However, the possibilities for using biomass in Germany are limited. Solar-
thermal energy, geothermic energy and heating pumps could also play a bigger role

82
4.3 Options for Protecting the Climate

8000
district heating
electricity

primary energy demand for heat in PJ


7000 f ossil f uels
hydrogen
6000 geothermal
heat pumps
solar thermal
5000
biomass

4000

3000

2000

1000

0
1990

1995

2000

2005

2010

2015

2020

2025

2030

2035

2040

2045

2050
Figure 4.12 Possible development of share of renewable energies in primary energy
requirement of the heat supply sector in Germany.

in the heat supply. In the long term hydrogen produced by renewable energies could
be used to cover remaining requirements. If, as explained earlier, direct electricity
and district heating could then be sourced from renewable electricity power plants,
Germany would have a carbon-free heat supply by 2050.

4.3.4 Increasing Efficiency and New Concepts for Traffic


At present, the biggest challenge is introducing climate-compatible changes in the
traffic sector. German car manufacturers have stubbornly resisted implementing
measures to protect the climate. Some car bosses even claim that the savings being
demanded would simply be physically impossible to implement. Yet, technically,
modern cars are by no means energy-saving marvels. Even modern combustion
engines only achieve 25% of average efficiency in an optimal case. At least 75%
of the energy content dissipates unused as waste heat into the environment. If the
Germans’ strong attachment to powerful luxury cars could be taken out of the equa-
tion, there is a technical certainty that energy use could be halved.
New and innovative driving concepts, such as electric cars and fuel-cell cars, can
increase efficiency levels, reduce energy use and, with renewable energies, slash
carbon dioxide emissions to zero (Figure 4.13).
In the short term biomass fuel could cover some of the fuel requirements of con-
ventional vehicles with combustion engines. However, the agricultural land avail-
able in Germany does not even come close to being able to provide enough biomass
fuel to cover the entire requirement. And, in addition, some of the available biomass
is needed to produce electricity and heat.
Due to considerable improvements in battery technology in recent years, electric
cars will soon be ready for production. During this transitional period they are still

83
4 Carbon-Free Energy – Vision or Utopia?

3000

_ PJ
electricity

primary energy demand for transport in


hydrogen
2500
biomass
f ossil f uels
2000

1500

1000

500

0
1990

1995

2000

2005

2010

2015

2020

2025

2030

2035

2040

2045

2050
Figure 4.13 Possible development of the share of renewable energy in primary energy
requirement of the transport sector in Germany.

partly being charged and driven with electricity from fossil power plants. However,
carbon dioxide emissions from these vehicles will plummet as the proportion of
renewable energies used to generate electricity increases.
Fuel-cell vehicles fuelled with renewable hydrogen will not come to the market until
some time in the future. Although considerable progress is being made in fuel-cell
technology, the structures for carbon-free production and the distribution of reason-
ably priced hydrogen are not yet in place. However, hydrogen technology will
eventually play a role in protecting the climate. It could also help to ensure that the
energy required in the road transport sector is carbon-free by the year 2050.

4.4 Reliable Supply Using Renewable Energies

Mathematically, on a yearly average renewable energies could cover the entire


energy requirement. However, many people still question how electricity supply can
be guaranteed – for example, if there is no wind blowing after the sun goes down.
In this case, none of the photovoltaic and wind power plants would be able to gener-
ate electricity.
Renewable energies can guarantee electricity supply even when fluctuations occur
provided a balanced mix of different renewable power plants is combined intelli-
gently through a central control system (Figure 4.14). To prove this, various com-
panies from the renewable energy sector initiated a combination power plant project
in 2007.
This combination plant linked and controlled 36 wind, solar, biomass and hydro-
power plants spread out all over Germany. It satisfied exactly one ten-thousandth

84
4.4 Reliable Supply Using Renewable Energies

Figure 4.14 Principle of coordinating combined power plants to guarantee renewable


electricity supply.

date: 24.07.06 25.07.06 26.07.06 27.07.06 28.07.06 29.07.06 30.07.06 25. July 17:00
time: 00:00
MW one week

8.0

7.0

6.0 solar energy


wind energy
5.0
biogas energy
storage energy
4.0 (pumped hydro)
export/import
3.0 electricity demand

2.0

1.0

Figure 4.15 Proportion of different renewable power plants covering energy requirements
during the course of a summer week with the combination power plant project. Source: www.
kombikraftwerk.de.

of the actual German electricity requirement. To do so, a central control centre


received information on load profiles as well as on current weather projections. If
wind and solar plants do not create enough electricity, then other plants have to
step in. In the case of a combination power plant, these would be biogas power
plants and a pumped storage station. As biogas is easily stored, electricity can be
generated at any time. Surplus supplies can then be stored temporarily in pumped
storage stations or used to charge batteries for electric cars. There are very few cases
when the production of wind and solar energy plants has to be cut back due to long

85
4 Carbon-Free Energy – Vision or Utopia?

sunny and windy periods. During these hours a small proportion of the surplus
is lost.
The results of the combination power plant project were very promising. They
showed that on a small scale optimal coverage of most energy needs is possible
(Figure 4.15). The desired reliability of energy supply was achieved. Therefore,
there is no convincing argument why renewable energies cannot be used to cover
all energy supply requirements using this approach in the near future.

86
5
5 Photovoltaics – Energy
from Sand

The word ‘photovoltaics’ comes from the two words ‘photo’ and ‘volta.’ In this
case photo stands for light and comes from the Greek phõs, or photós. The Italian
physicist Alessandro Giuseppe Antonio Anastasio Count Volta, who was born in
1745, was the inventor of the battery and together with Luigi Galvani is considered
the discoverer of electricity. There is not much that associates him with photovolta-
ics. However, in 1897, 70 years after Volta’s death, the measurement unit for electric
current was named volt in his honour. Photovoltaics therefore stands for the direct
conversion of sunlight into electricity.
While fiddling with electro-chemical batteries with zinc and platinum electrodes,
the nineteen-year-old Frenchman Alexandre Edmond Becquerel found that the elec-
tric voltage increased when he shone a light on them. In 1876 this phenomenon was
also proven with semiconductor selenium. In 1883 the American Charles Fritts
produced a selenium solar cell. Due to the high prices of selenium and manufactur-
ing difficulties, this cell was not in the end used to produce electricity. The physical
reason why certain materials produce electric voltage when radiated with sunlight
was not understood at the time. It was not until many years later that Albert Einstein
was able to specify the photo effect that causes this. He eventually received the
Nobel Prize for this work in 1921.
The age of semiconductor technology began in the mid-1950s. The semiconductor
material silicon, which often occurs naturally, became all the rage in technology,
and in 1954 the first silicon solar cell finally made its appearance at American Bell
Laboratories. This was the basis for the successful and commercial further develop-
ment of photovoltaics.

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

87
5 Photovoltaics – Energy from Sand

5.1 Structure and Function

5.1.1 Electrons, Holes and Space-Charge Regions


Understanding the relatively complicated way that solar cells work requires immer-
sion into the most extreme depths of high physics. The small applied model shown
in Figure 5.1 explains roughly the principle involved. There are two horizontal
levels. The second level is located a bit higher than the first one. The first level has
a large number of small hollows filled to the top with water. The water here cannot
move by itself. Now someone starts to throw small rubber balls at the first level. If
a ball hits a hole, the water splashes upwards and ends up on the second level. Here
there are no hollows to contain the water. The second level is therefore inclined so
that the water runs off and reaches the draining groove on its own. This groove is
connected to the second level through a pipe and as the water flows through, it drives
a small waterwheel with a dynamo. When the water reaches the lower level, it fills
up the hollows again. The cycle can start all over again with new rubber balls.

Figure 5.1 Model illustrating the processes of a solar cell.

However, we want to use solar cells not to produce a water cycle but to generate
electric current to run electrical appliances. Electric current is created from the flow
of negative-charge carriers, called electrons. These are the same as the water in our
simple model. The solar cell needs a material in which two levels can be found: one
level in which the electrons are firmly affixed like the water collecting in the
hollows, and a second level where the electrons are able to move freely.
Semiconductor materials normally have precisely these properties. Tiny particles of
light, called photons in physics, correspond to the rubber balls and are able to raise
the electrons to the second level.

Conductors, Non-conductors and Semiconductors


Conductors such as copper always conduct electric current relatively well but
non-conductors such as plastics conduct almost no electricity at all. In con-
trast, semiconductors – as the name indicates – only conduct electric current sometimes,
for example at high temperatures, when fed with electric voltage or when radiated with
light. These effects are used in the production of electronic switches like transistors,
computer chips, special sensors and even solar cells.

88
5.1 Structure and Function

Organic semiconductors are available in addition to elementary semiconductors, such as


silicon (Si), and compound semiconductors, such as gallium arsenide (GaAs), Cadmium
telluride (CdTe) and copper indium diselenide (CuInSe2). All these materials are used
in photovoltaics.

The tilt in our simple model is important because it enables the cycle to function
perfectly. Otherwise the water will not collect on its own in the rain gutter. With
semiconductors the second level must also have an incline that enables the electrons
to gather on one side. In contrast to our simple model, it is not gravity that is used
to collect the electrons but instead an electric field, which pulls the negatively
charged electrons to one side. In order to produce this field, a semiconductor must
be ‘doped’. One side of the semiconductor is deliberately contaminated with an
element like boron and the other side with a different element like phosphorus. As
boron and phosphorus also have a varying number of electrons, they produce the
necessary incline. The crossing area is called a space-charge region. An electric field
is created here, which pulls the electrons to one side. There external contacts collect
them and they flow back to the first level through an external electric circuit. In the
process they produce electrical energy.
Figure 5.2 shows the principal structure of a silicon solar cell. In technical jargon
the different doped sides of the silicon wafer are called n-doped and p-doped silicon,
respectively. Between the two areas is a barrier layer with the space-charge region.
Light in the form of photons separates negatively charged particles (electrons) and
positively charged particles (holes) and ensures that the electrons are able to move
about freely at the second level. In contrast to the simple model shown, the holes
also move. The electrons and holes are separated by the space-charge region. Thin
front contacts collect the electrons on the front side of the cell.

solar radiation (photons)

front contact
consumer
n-doped
silicon

electron
-
+ - hole
+
+ - - +

- boundary layer
+
p-doped
back contact silicon

Figure 5.2 Structure and processes of a solar cell (Quaschning, 2007).

89
5 Photovoltaics – Energy from Sand

However, not every light particle ensures that an electron is separated from the hole.
If the energy of the photon is too low, the electron will fall back into the hole. On
the other hand, if the energy of the photon is too high, only part of it is used to
separate the electron from the hole. Some photons also move through the solar cell
unused; others are reflected by the front contacts.

5.1.2 Efficiency, Characteristics and MPP


The efficiency of a solar cell specifies which part of the solar radiant power is con-
verting the cell into electric power.

Solar Cell Efficiency

Pel ⎛ electrical power output ⎞


η= ⎜⎝ efficiency = ⎟
Φ incoming solar radiant power ⎠

The higher the efficiency, the more electric power the solar cell can generate per
square metre. In addition to the type of materials used, the quality of the manufac-
turing also plays a major role. Today silicon cells in mass production reach a
maximum efficiency of over 20%. Close to 25% efficiency has already been reached
in laboratories (Table 5.1).

Table 5.1 Efficiency of different solar cells.

Cell material Max. cell Max. cell Typical Surface


efficiency efficiency module req. for
(lab) (mass prod.) efficiency 1 kWp
Monocrystalline silicon 24.7% 22.0% 15% 6.7 m2
Polycrystalline silicon 20.3% 17.4% 14% 7.2 m2
Amorphous silicon 12.1% 6.8% 6% 16.7 m2
CIS/CIGS 20.0% 11.6% 10% 10.0 m2
CdTe 16.5% 12.0% 7% 14.3 m2
Concentrator cells 41.1% 36.5% 28% 3.6 m2

Incidentally, conventional petrol engines do not reach a higher level of efficiency


than silicon cells. Compared to the 5% efficiency of the first solar cells in 1954,
technology has advanced considerably. If individual solar cells are packaged into
photovoltaic modules, the efficiency drops somewhat due to the space necessary
between the cells and the module frames. It is hoped that other materials can be
used in future to produce further cost savings. In comparison with silicon cells, more

90
5.1 Structure and Function

improvements in efficiency are needed in this area. Concentrator cells in which


sunlight is concentrated through mirrors or lenses reach a very high efficiency but
are considerably more expensive than normal silicon cells.
In addition to efficiency, there are other parameters that describe photovoltaic
modules. The current-voltage characteristic is usually found in relevant data sheets.
The maximum current ISC flows with a short-circuited photovoltaic module. The
short-circuiting is not dangerous to the module. The short-circuit current is limited
and depends on the solar radiation intensity. If nothing is connected to the photo-
voltaic module, it is in open circuit and no electricity flows. It then adjusts itself to
the open-circuit voltage VOC. A photovoltaic module is unable to produce any power
when in short circuit or open circuit. Between open circuit and short circuit the
current depends on the voltage. The principal process of the curve is similar for all
solar modules (Figure 5.3 and Table 5.2).

ISC maximum power


point MPP
IMPP
photovoltaic current I

photovoltaic power
PMPP = VMPP · IMPP

photovoltaic voltage V VMPP VOC

Figure 5.3 Current-voltage curve for a photovoltaic module.

In practice, the aim is to take maximum power from the photovoltaic module. This
corresponds to the largest rectangle that can be slid under the curve. The top edge
on the right of the rectangle on the curve is called the MPP (maximum power point).
The voltage that belongs to the MPP is called MPP voltage, in short VMPP. A pho-
tovoltaic module provides maximum power with this voltage. In practice, operation
close to that of the MPP can be achieved, for example when a battery with voltage
close to that of the MMP voltage is connected or where an inverter automatically
adjusts the MMP voltage on the photovoltaic module.
The current of photovoltaic modules, and hence the power, drops according to the
number of incoming photons, and thus the radiation intensity of the sunlight. If the
solar radiation intensity is halved, then the power of the photovoltaic modules is
also reduced by half. The power of photovoltaic modules also drops with high
temperatures. If the temperature rises by 25 °C, the power of the crystalline solar
cells falls by about 10%. Therefore, when photovoltaic modules are installed, care
should be taken that they are always well ventilated and a draught cools the modules
(Table 5.2).

91
5 Photovoltaics – Energy from Sand

Table 5.2 Important parameters for photovoltaic modules.

Parameter Symbol Unit Description


Open-circuit VOC Volt, V Voltage of PV module in open-circuit
voltage operation without connected load
Short-circuit ISC Ampere, A Current of PV module in short-circuit
current operation with short-circuited module
MPP voltage VMPP Volt, V Voltage when a photovoltaic module
produces maximum power
MPP current IMPP Ampere, A Current belonging to MPP voltage
MPP power PMPP Watt, W Maximum power a PV module can produce

Standard test conditions (STC) for the comparison of photovoltaic modules have
been agreed internationally. Accordingly, the MPP power of solar cells and modules
is determined on the basis of solar radiation intensity of 1000 watts per square metre
and a module temperature of 25 °C. As radiation intensity is usually lower in practice
and photovoltaic modules can warm up to 60 °C in the summer, the MPP power
determined on the STC basis represents a maximum value. This value is achieved
in very few cases and seldom exceeded. Therefore, this power also has the unit watt
peak, abbreviated Wp.

5.2 Production of Solar Cells – from Sand to Cell

5.2.1 Silicon Solar Cells – Electricity from Sand


Silicon, the raw material for computer chips and solar cells, is the second most
common element in the earth’s crust after oxygen. However, in nature silicon occurs
almost exclusively as an inclusion in quartz, sand and silicate rock or as silicic acid
in the world’s oceans. Even the human body contains around 20 milligrams of
silicon per kilogram of body weight.
Pure silicon is usually extracted from quartz sand. Chemically, quartz sand is pure
silicon dioxide (SiO2). For silicon to be produced from it, high temperatures are
used to sequester the oxygen atoms (O2). This process is called reduction and is
carried out in arc furnaces at temperatures of around 2000 °C. The result is industrial
raw silicon with a purity of 98 to 99%.
Raw silicon has to be purified further before it can be used to produce solar
cells. The Siemens method is the normal procedure followed. Hydrogen chloride is
used to convert the raw silicon into trichlorosilane, which is then distilled. At high
temperatures of 1000 to 1200 °C the silicon is then separated again into long
bars. The polycrystalline solar silicon produced in this way has 99.99% purity
(Figure 5.4).

92
5.2 Production of Solar Cells – from Sand to Cell

Figure 5.4 Polycrystalline silicon for solar cells. Left: Raw silicon. Middle: Silicon blocks. Right: Silicon
wafers. Photos: PV Crystalox Solar plc.

The silicon is melted down again to produce semiconductor silicon for computer
chips and monocrystalline solar cells. In the crucible process invented by Polish
chemist Jan Czochralski, a silicon crystal is dipped into a crucible with a silicon
melt and then slowly pulled upwards in a rotating movement. The melted silicon
attaches itself to the crystal and a long, round silicon rod is created. In the process
the silicon crystals align in one direction. This creates monocrystalline silicon. Most
of the impurities remain in the melt crucible so that the semiconductor silicon is left
with purities of over 99.9999%.
In the next step, band saws cut the long silicon rods into thin slices, called wafers.
This sawing process produces major waste, and up to 50% of the valuable silicon
material is lost as a result. The alternative is for two thin wires to be pulled through
the liquid silicon melt. With this procedure, thin silicon wafers are formed between
the two wires. Immersion in acid will remove sawing damage from wafers and
smooth the surfaces. Several years ago silicon wafers had a thickness of 0.3 to
0.4 mm. To save material and costs, attempts are now being made to reduce the
wafer thickness to under 0.2 mm. Technically, this is a major challenge as the ultra-
thin wafers must not break apart.
The finished wafers are exposed to gaseous doping materials. This produces the
p- and n-layers described earlier. A transparent anti-reflection layer of silicon nitride
less than a millionth of a millimetre thick gives the silicon solar cell its typical blue
colour. This layer reduces the reflection loss of the silver-grey silicon on the top
side of the solar cell. The darker the cell appears, the less light the cell is reflecting
(Figure 5.5).
The front and back contacts are then applied using screen printing. To reduce the
losses to the opaque front contacts, some manufacturers conceal them under the
surface or try to move them so that they are also on the back of a cell. Although

93
5 Photovoltaics – Energy from Sand

Figure 5.5 Polycrystalline solar cells with anti-reflection layers before the front contacts
are applied. Photo: BSW, www.sunways.de.

this increases the efficiency of a cell, it is also a more complicated and expensive
way to manufacture cells. The finished cells are then finally tested and sorted accord-
ing to performance class for further processing into photovoltaic modules.

5.2.2 From Cell to Module


Silicon solar cells are usually square in shape. The length of the edge is measured
in inches. Originally solar cells were typically 4 inches (approx. 10 cm) in length.
In the meantime a measurement of 6 inches (approx. 15 cm) has established itself
as the standard. Some manufacturers are already producing 8-inch (approx. 20 cm)
solar cells. Large solar cells require fewer processing steps to be made into modules.
However, the risk that the cells will break during further processing is also greater.
The current increases with the size of a solar cell, whereas the voltage remains
constant. The electric voltage of a solar cell is only 0.6 to 0.7 volts.
Practical applications clearly require high voltages. Therefore, many cells are inter-
connected in series to form solar modules. The front contacts of a cell are connected
to the back contacts of the next cell using soldered-on wires. It takes 32 to 40 cells
connected in series to produce a sufficiently high voltage to charge 12-volt batteries.
Higher voltages are needed to feed into the grid through inverters. Solar modules
with an even higher number of cells interconnected in series are available for this
purpose.
Solar cells are very sensitive, break easily and corrode when in contact with mois-
ture, so they must be protected. Consequently, they are imbedded in a layer of
special plastic between the front glass panel and a plastic film on the back (Figure
5.6). Some manufacturers also use glass for the back. The glass provides mechanical
stability and must be very translucent. The plastic material used for imbedding the
solar cells consists of two thin films of ethylene-vinyl-acetate (EVA). At tempera-
tures of around 100 °C these films bond with the cells and the glass. This process
is called lamination. The finished laminate then protects the cells from the effects
of the weather, especially moisture.

94
5.2 Production of Solar Cells – from Sand to Cell

Figure 5.6 Structure of a photovoltaic module.

The connections of the solar cells are linked to a module connection box. Individual
faulty cells or uneven shade can damage a PV module. Bypass diodes that compen-
sate for affected cells when a fault occurs are designed to prevent any damage. These
diodes are also usually integrated into the module junction boxes.

5.2.3 Thin Film Solar Cells


Crystalline solar cells require a comparatively large amount of costly semiconductor
material. Different production methods using thin film cells are being tested to
reduce the amount of material needed. Whereas crystalline solar cells reach thick-
nesses in the order of tenths of millimetres, thin film solar cells are thousandths of
a millimetre thick. The production principle is similar even when different materials
such as amorphous silicon (a-Si), cadmium telluride (CdTe) or copper indium dis-
elenide (CIS) are used.
The base of thin film solar cells is a substrate that is usually made of glass. Plastic
can be used instead of glass for the substrate to produce modules that are flexible
and bendable. A thin TCO (Transparent Conductive Oxide) layer is applied to the
substrate using a spraying technique. A laser or a micromiller then separates this
layer into strips. The individual strips constitute the single cells within the later solar
module. Like crystalline cells, these cells are also contacted in such a way that
they are connected in series to increase the electric voltage. The long strips make
it visually easy to distinguish thin film modules from crystalline solar modules
(Figure 5.7).
The semiconductor and doping materials are then vaporized at high temperatures.
When silicon is vaporized as a semiconductor material, the crystalline structure of
the silicon is lost. This is then referred to as amorphous silicon. A screen-printing
procedure then applies materials like aluminium to the back side contact. A layer
of polymer seals the cell at the back to protect it from moisture.
The efficiency of thin film modules is currently still considerably less than that of
crystalline photovoltaic modules. This means that a larger surface is required for

95
5 Photovoltaics – Energy from Sand

Figure 5.7 Cross-section of a thin film photovoltaic module.

the same power output, and, therefore, more assembly is required and the associated
costs are higher. However, because less material is used, in the long term the cost
of thin film modules could fall below that of crystalline modules. If the efficiency
of thin film modules can also be increased, this technology could put an end to the
current predominance of crystalline silicon solar cells.
In addition to thin film materials, other technologies are currently being tested.
Pigment cells and organic solar cells could eventually offer a cost-effective alterna-
tive to current technologies. At the moment it is almost impossible to project which
technologies will find acceptance in 30 or 40 years. But competition is bringing new
life to the sector. Costs will continue to drop due to the competition of different
technologies for photovoltaics. So in the medium term photovoltaics will be very
important in the effort to provide low-cost and climate-compatible electricity supply.

5.3 Photovoltaic Systems – Networks and Islands

5.3.1 Sun Islands


With photovoltaic systems a distinction is made between island systems and grid-
coupled systems. Solar island systems work autonomously without being connected
to an electricity grid. For example, they are often used in small applications, such
as wristwatches and pocket calculators, because in the long run they are less expen-
sive than energy supplied by disposable batteries, and a network cable in this case
would be highly impractical. Solar systems are also popularly used for small systems
like car park ticket machines. In this case it is less expensive to install a photovoltaic
system than to lay network cables and install a meter (Figure 5.8).
The big market for solar island systems is in areas that are far from an electricity
grid. Globally, around two billion people have no access to electricity. Even in

96
5.3 Photovoltaic Systems – Networks and Islands

Figure 5.8 Photovoltaic island systems offer advantages for many applications compared to grid
connections.

industrialized countries there are towns that are very remote from the grid and where
any kind of cabling would be extremely expensive. There are alternatives to using
island networks to supply solar energy – such as diesel generators. A solar system
often compares favourably in terms of cost and supply reliability, specifically in
cases where demand for electricity is low. Nevertheless, the costs of photovoltaic
island systems are still relatively high. Therefore, they are currently used mostly for
small output requirements. For cost reasons, diesel generators and grid connections
are usually given preference when a high demand for electricity exists.
Solar island systems are comparatively straightforward and can even be installed by
people with limited technical skills (Figure 5.9). A battery ensures that supply is
available at night or during periods of bad weather. For reasons of cost, lead batteries
are normally used. In principle, 12-volt car batteries are also an option. Special-
purpose solar batteries have a considerably longer lifetime but are also more expen-
sive. As batteries can quickly be ruined as a result of leakage or overcharging, a
charge regulator protects the battery. The battery, the power consumer and the photo-
voltaic module are connected directly to the charge regulator. It is important to make
sure the plus and minus poles are not switched as this will cause a short circuit.
When the battery is nearly empty, the power consumer is switched off. Although
the lack of electricity is a nuisance, this is better than a defective battery. As

97
5 Photovoltaics – Energy from Sand

Figure 5.9 Principle of a photovoltaic island system.

soon as the battery reaches a certain charge level again, the charge regulator auto-
matically switches the power consumer back on. Once the battery is fully charged,
the charge regulator separates the photovoltaic module and prevents the battery from
overloading.
The costs can be kept low if the power consumers are economical users. Battery
voltage higher than 12 volts is recommended if power consumers have high require-
ments, as otherwise the losses in the lines will be too high. Island systems work on
a direct voltage basis, so, if possible, only direct voltage power consumers should
be utilized. Special refrigerators, lamps and even entertainment electronics are being
offered with 12 or 24 direct voltage supply. If alternating voltage power consumers
are operated, an island inverter first has to convert the direct voltage of the battery
to alternating voltage.
Normal low voltages are usually relatively safe, at least as far as contact with the
voltage is concerned. As batteries are pure bundles of power, improper handling
can cause short circuits, fire or even explosions. Battery rooms should always have
good ventilation as hydrogen gas can build up in them. Lead batteries contain diluted
acid. As time goes by, water evaporates from the battery and the battery has to be
refilled regularly. With maintenance-free batteries the water is bound in a gel and
cannot escape.
A large array of photovoltaic modules can end up costing the same as a mid-range
car, and even individual photovoltaic modules are quite expensive, so thieves are
making life increasingly difficult for the operators of remote photovoltaic systems.
Solar modules installed on remote roads are particularly vulnerable to theft. As
island systems are usually installed in even more out-of-the-way places, the risk of
theft is especially high. This risk should be minimized in the installation. Optimally,
photovoltaic systems should not be visible from public roads. However, if this
cannot be avoided, systems should at least be assembled in places that are difficult
to access (Figure 5.10).

98
5.3 Photovoltaic Systems – Networks and Islands

Figure 5.10 Typical places for using photovoltaic island systems. Left: Electricity supply to a village in
Uganda. Right: Mountain hut. Source: SMA Technologie AG.

Small solar-operated systems can only contribute minimally, if even at all, to climate
protection. On the other hand, photovoltaic-operated pocket calculators and watches
help reduce the mountain of small spent batteries.

5.3.2 Sun in the Grid


Photovoltaic systems that feed into a public grid are built in a different way from
island systems. First of all, they normally require a larger number of modules. With
crystalline solar modules a top output of 3.5 to 5 kilowatts peak (kWp) can be
installed for a 25 square metre area. In Germany this generates between 3000 and
4000 kilowatt hours with the right kind of roof; in Southern Europe or better sites
in the USA the figure increases to around 50%. This easily covers the electricity
requirements of an average German household. An ideal surface can usually be
found quite easily on the roofs of single-family homes (Figure 5.11).
As solar modules produce direct voltage but the public grid works with alternating
voltage, an inverter is also required. An inverter converts the direct voltage of a
photovoltaic module into alternating voltage. Modern inverters have high demands
placed on them. They are required to have a high level of efficiency to ensure that
only very little valuable solar energy is lost during the conversion into alternating
voltage. Modern photovoltaic inverters achieve efficiency of close to 95% or even
more. What is important is that efficiency is high even with a partial load, such as
in cloudy weather. European efficiency describes average inverter efficiency based
on Central European climate conditions, while CEC efficiency is for Californian
conditions.
Inverters also have to monitor the grid constantly and switch off the solar feed when
a general grid outage occurs. If the electric company wants to do work on the grid
and needs to shut off the electricity to do this, the solar current still being fed in
could harm the workers. If a public power outage occurs, owners of photovoltaic
systems will unfortunately end up sitting in the dark even though there is nothing

99
5 Photovoltaics – Energy from Sand

Figure 5.11 Principle of grid-connected photovoltaic system.

wrong with their solar system. Technically, a solar system can still be kept running
in island operation if the grid fails. However, as power failures are very infrequent,
hardly anyone makes the additional technical provisions that would be necessary.
An inverter does not only convert voltage. It also ensures that photovoltaic modules
are working with optimal voltage and delivering the maximum possible power. The
setting of optimal voltage is also called MPP tracking. When a system is planned,
it is important that the number of photovoltaic modules is coordinated with the
inverter. The leading manufacturers of inverters usually offer relevant design pro-
grammes free of charge.
Shadowing can also cause problems. Photovoltaic systems react sensitively and
performance suffers even if only part of a system is in the shade. If two cells are in
shade, the entire photovoltaic module can stop working. Therefore, a site that has
minimal shade is more important for the installation of a system than an optimal
orientation towards the sun.
Large grid-linked photovoltaic systems feed all generated electricity into the public
grid. They function as pure solar power plants. With large systems it can be benefi-
cial to track them towards the sun (Figure 5.12). Systems that use tracking can
increase power output by 30% on an annual average. However, tracking also
increases the investment cost and requires additional maintenance because of the
mechanical parts required. Therefore, tracking pays particularly in situations where
large, well-maintained systems are installed.
The neatest looking installations of photovoltaic systems are those on roofs and
façades (Figures 5.13 and 5.14). Less building material is also needed for this type
of installation, which is a cost advantage as these systems are still very expensive.
Furthermore, compared to expensive marble façades, photovoltaics can sometimes
turn out to be a real bargain.

100
5.3 Photovoltaic Systems – Networks and Islands

Figure 5.12 The ‘Gut Erlasee’ tracked solar power plant in Bavaria, Germany, has a
total output of 12 megawatts. Source: SOLON AG, Photo: paul-langrock.de.

Figure 5.13 Photovoltaic facade system. Photo: SunTechnics.

Some of the solar power can be used directly in the building where the system is
installed. If the photovoltaic system produces more power than is needed, it feeds
the surplus electricity into the public grid. However, if the output of the solar system
is not sufficient to cover the building’s own requirements, then the rest of the elec-
tricity that is needed is drawn from the grid. In a sense, the grid is like a storage
unit. But the grid is not able to store electricity. When solar power is fed into a grid,
then other power plants cut their production. As a result, solar systems reduce the
emissions of existing power plants. If there is insufficient output, this then has to

101
5 Photovoltaics – Energy from Sand

Figure 5.14 Photovoltaic systems on single-family homes. Foto: SunTechnics.

be sourced from other power plants. These power plants do not necessarily have to
be coal-driven, gas or nuclear plants. On the contrary, photovoltaic systems work
compatibly with other renewable energy plants, such as wind power, hydropower
and biomass plants.
An extra electricity meter is normally required to calculate the photovoltaic power.
This meter calculates the quantity of electric energy that has been fed into the
grid, which is then credited according to existing tariffs. If, as in Germany, the
feed-in tariff for solar energy is higher than the normal tariff for electricity, it can
make sense to feed all the electricity produced by a photovoltaic system into the
grid through a feed-in meter and then to buy back some of the supply at the lower
tariff. In other countries solar power is sometimes credited at a noticeably lower
rate than normal electricity. If this is the case, it makes more sense to use the
electricity generated through solar power first and only feed excess supplies into
the grid.
Connecting to the grid in Germany is usually not a problem. An electrician is needed
to connect the system and the relevant electricity supplier has to be notified. In
addition to the connection protocol, technical documents relating to the photovoltaic
system are submitted with the application. It is important that the system complies
with the general regulations, and this is not generally a problem with current system
suppliers. In most cases, a representative of the electricity supplier inspects the
system. The reimbursement is handled automatically and is calculated according to
the meter reading. The electric company and the operator of the photovoltaic system
usually also sign a contract. However, this is not always a stipulation.

102
5.4 Planning and Design

5.4 Planning and Design

5.4.1 Planned on the Grid


The first thing to do when planning a solar system is to check whether a system can
even be installed. From the standpoint of planning laws, solar systems are structural
items – even if they are simply screwed to the roof of a house. The authority respon-
sible for local building regulations determines whether approval is necessary, and,
if so, what type of approval. In most cases photovoltaic systems do not need
approval from the local authority unless they are erected in the middle of a green
field. Complications arise when the sites for such systems are subject to architectural
conservation laws. A permit from the relevant authority is required if a solar system
is to be installed on top of or near a protected building. In addition to obtaining
planning permission, it is always a good idea to consult the building regulations
and the development plans of the local authority. These plans will include any
conditions that the local authority has stipulated for the construction of photovoltaic
systems.
If there are no legal obstacles, the planning phase can start. If a photovoltaic system
is to be installed on the roof of a house, it first needs to be determined which parts
of the roof can be used for a solar-thermal system (see Chapter 6). As photovoltaic
systems are sensitive to shade, it is recommended that the part of the roof used to
generate solar power is shade-free. The installation should not be placed near chim-
neys, aerials and other roof structures.

Installable Photovoltaic Power

Once the type of solar module has been selected, an approximate calculation can
be made of the installable photovoltaic power based on remaining roof area A and
efficiency η (compare Table 5.1):
PMPP = A ⋅ η ⋅ 1 kW
m2
.

The installable power on a useable area of 28.6 m2 with a module efficiency of 14%
(0.14) is
PMPP = 28.6 m 2 ⋅ 0.14 ⋅1 kW
m2
= 4 kWp .

The annual system yield can be calculated on the basis of this power output. What
first needs to be determined is the supply of solar energy available. The map in
Figure 5.15 shows the yearly total for solar radiation in Europe based on averages
over many years. The radiation data for the USA can be found under http://www.

103
5 Photovoltaics – Energy from Sand

nrel.gov/rredc/. Fluctuations of over 10% in solar supply are possible between the
individual years given.

10°W 5°W 0° 5°E 10°E 15°E 20°E 25°E


60°N

60°N
2
kWh/m
> 2420
2381−2420
2341−2380
2301−2340
2261−2300
55°N

55°N
2221−2260
2181−2220
2141−2180
2101−2140
2061−2100
2021−2060
1981−2020
1941−1980
1901−1940
1861−1900
1821−1860
1781−1820
50°N

50°N
1741−1780
1701−1740
1661−1700
1621−1660
1581−1620
1541−1580
1501−1540
1461−1500
1421−1460
1381−1420
1341−1380
1301−1340
45°N

45°N
1261−1300
1221−1260
1181−1220
1141−1180
1101−1140
1061−1100
1021−1060
981−1020
941−980
901−940
861−900
821−860
40°N

40°N
781−820
741−780
701−740
661−700
621−660
581−620
541−580
501−540
<501
35°N

35°N
10°W 5°W 0° 5°E 10°E 15°E 20°E 25°E

Figure 5.15 Average yearly totals for solar radiation energy in Europe. Source: Meteonorm, www.
meteonorm.com.

However, the values only apply to the horizontal orientation of a system. If a system
is mounted on a sloping roof, the roof determines the orientation of the system. In
an optimal case, the roof would be tilted about 30° to the south in Europe and North
America. With an optimal orientation towards the sun the supply of solar radiation
increases by around 10%. However, good radiation values are still achievable even
if the orientation is less favourable. Figure 5.16 shows the tilt gains and losses for
all possible orientations for Berlin. These values can also be applied to other loca-
tions in Central Europe.
If a photovoltaic system is to be installed on a roof or in a green field, the photo-
voltaic modules can be orientated optimally 30° towards the south. If the solar
modules are set up in several rows one behind the other, they will create shade for
one another as the sun goes down. As a result, a distance at least twice the module
height should be maintained between the rows of modules. In this case, however,
only one-third of the area is useable. The loss due to shade then usually amounts to
less than 5%.

104
5.4 Planning and Design

110 105 100 95 90 85 80 75 70 65 60 55 50


%

90°
vertical

80°

70°

angle of inclination
60°

50°

40°

30°

20°

10°
horizontal

-180° -150° -120° -90° -60° -30° 0° 30° 60° 90° 120° 150° 180°
north east SE south SW west north
orientation

Figure 5.16 Change in annual solar radiation in Berlin depending on orientation and tilt
angle of photovoltaic systems.

The proportion of solar radiation that photovoltaic modules convert into electric
energy depends on the quality of the system. Solar modules rarely achieve the rated
power indicated. Dust, bird droppings, increases in air temperature, line losses,
reflection, inverter losses and other factors reduce performance. The relationship
between real efficiency and rated efficiency is called performance ratio (PR). Table
5.3 lists the criteria for performance-ratio values for grid-connected photovoltaic
systems.

Table 5.3 Performance ratio for grid-connected photovoltaic systems.


Performance Description
ratio PR
0.85 Top-quality system, very good ventilation, no shade, minimal pollution
0.8 Very good system, good ventilation, no shade
0.75 Average system
0.7 Average system with minimal losses due to shade or poor ventilation
0.6 Poor system with major losses due to shade, pollution or system failures
0.5 Very poor system with large areas of shade or defects

105
5 Photovoltaics – Energy from Sand

Electric Energy Yield of Photovoltaic systems

A rough calculation of the quantity of energy fed in annually by a grid-connected


photovoltaic system can be made on the basis of the annual solar radiation Hsolar in
kWh/(m2 a) from Figure 5.15, the losses or gains through orientation and tilt angle
forientation (Figure 5.16), the rated or MPP power PMPP of photovoltaic modules in kWp
and the performance ratio PR (compare Table 5.3):
H solar ⋅ forientation ⋅ PMPP ⋅ PR
E electric = .
1 kW
m2

The yield of a photovoltaic system in Berlin that is orientated approximately 20°


towards the south-southeast and tilted approximately 20° towards the south-south-
west and has an MMP power of 4 kWp will be used as a sample calculation. The
annual solar radiation according to Figure 5.15 is around 1020 kWh/(m2 a). According
to Figure 5.16, the tilt gains are forientation = 110% = 1.1 based on the south orienta-
tion. Based on the performance ratio of an average system PR = 0.75, the annual
quantity of fed-in electric energy is:
1020 kWh
m2 a
⋅1.1⋅ 4 kWp ⋅ 0, 75 kWh
E electric = = 3366 .
1 kW
m2
a

This corresponds approximately to the consumption of an average German house-


hold. Therefore, as a yearly average about 30 m2 of roof area is necessary for a
household to cover its total electricity requirements using a solar system. The spe-
cific yield related to 1 kWp is often given for a system. In the example above, the
total is 825 kWh/(kWp a).

The rough calculation of yield described here naturally incorporates certain inac-
curacies. However, the order of magnitude of the system yield is correct. It can be
used in the following section to examine the economic viability. Internet tools and
computer programmes are available to conduct a more precise analysis. Professional
firms also offer to carry out calculations of this type. In any case, if a large system
is to be installed, it is advisable to get an expert to produce an appraisal on yield to
avoid unpleasant surprises later on. Banks also often require a copy of the relevant
appraisal.

■ http://www.valentin.de/onlineberechnung/pv PV online calculations


■ http://www.nrel.gov/rredc Renewable Resource Data
Centre
■ http://re.jrc.ec.europa.eu/pvgis/ Photovoltaic Geographical
Information System PVGIS

It is not necessary for a grid-connected system to be large enough to cover the total
electricity requirements of a household. Depending on the roof size and how much
money is available, the system can turn out to be larger or smaller than optimally

106
5.4 Planning and Design

needed. Many roofs have enough space for large photovoltaic systems that are
capable of supplying more electricity than required and can, therefore, be optimal
in protecting the climate. But even small systems can make a contribution.

From Thinking about Having a PV System on the Roof to


Getting a Quote

■ Determine orientation and tilt angle of roof.


Is the roof suitable from the point of view of orientation and tilt angle?
Recommendation: min. 100% according to Figure 5.16.
■ Establish shading conditions on the roof.
Does my roof have little shade?
Recommendation: Take shaded areas into account during planning; also pay
attention to chimneys, aerials and lightning conductors.
Identify available roof area with little shade.
■ Calculate installable MPP power for the photovoltaic system.
Recommendation: Efficiency min. 12% in formula above.
■ Calculate possible annual electric energy yield.
Use formula above or online tools.
■ Request quotations.

5.4.2 Planned Islands


The design of photovoltaic island systems is fundamentally different from grid-
connected systems. Island systems cannot rely on the public electricity grid if there
is no sun. Instead these systems use sufficiently large batteries to avoid any power
failures. However, in principle a battery is only used to bridge the days that are
less than optimal. Therefore, photovoltaic modules also have to deliver as high a
yield as possible during the months with the least sunshine. The recommendation
for reliable operation in winter is to place photovoltaic modules at a steeper angle
than is necessary with grid-connected systems, which are angled optimally for
operation all year round. In Europe and North America a tilt of around 60° to 70°
towards the south provides optimal solar yield in the month of December. The closer
one gets to the equator, the less marked are the differences between summer and
winter. In these parts of the world a flatter installation is also sufficient for winter
operation.
The objective of island systems is not to achieve as high an output yield as possible
but to provide reliable electricity supply to certain consumers. Consequently, the
main criterion for the design of such a system is the estimated consumption level
for the worst month of the year. In Europe and North America this is December.
Furthermore, solar systems are equipped with a certain safety margin. In a normal
case this should amount to an extra 50%. As with grid-coupled systems, the perform-
ance ratio PR takes into account the losses.

107
5 Photovoltaics – Energy from Sand

Table 5.4 Monthly and yearly totals of solar radiation in kWh/m2 for different locations and orientations.

Location Berlin Madrid Los Angeles Cairo


Orientation 30 ° south 60 ° south 30 ° south 60 ° south 30 ° south 30 ° south
February 28 31 100 113 139 157
April 117 106 172 141 201 204
June 151 127 214 163 175 214
August 153 135 209 175 223 220
October 68 87 105 137 164 201
December 21 25 73 82 128 146
Year 1139 1042 1855 1675 2089 2310

PV Power Needed for Island Systems

The required MPP power PMPP of a photovoltaic module can be roughly calculated
based on the solar radiation Hsolar,m during the worst month in kWh/m2 (Table 5.4),
the electricity requirement Edemand,m for the same month, a safety margin fS of at least
50% and the performance ratio PR (on average 0.7):
(1 + fS ) ⋅ Edemand,m 1 kW
PMPP = ⋅ m .
2

PR H solar,m
The battery should be dimensioned so that it is planned to be discharged to only
half its power and can supply the total electricity requirements for a number of
reserve days. In Central Europe up to five reserve days are enough to ensure reliable
operation in the winter; in countries that get more sun only two to three reserve days
would be enough. The required battery capacity is calculated on the basis of the
battery voltage Vbat (e.g. 12 volts):
2 ⋅ Edemand,m dr
C= ⋅
Vbat 31
For example, a photovoltaic system should be capable of operating an 11-watt low-
energy light bulb in a summerhouse for three hours every day in the winter. The monthly
electricity requirement for one month is then Edemand,m = 31 · 11 W · 3 h = 1023 Wh.
With a safety margin of 50% = 0.5 and a performance ratio of 0.7, a module in Berlin
orientated towards the south and tilted 60° then has a required MPP power of
(1 + 0.5) ⋅1023 Wh 1 kW
PMPP = ⋅ mkWh = 87.7 W.
2

0.7 25 m2
With four reserve days and a battery voltage of 12 V, the battery capacity is calcu-
lated as follows:
2 ⋅1023 Wh 4
C= ⋅ = 22 Ah.
12 V 31

108
5.5 Economics

5.5 Economics

Photovoltaics is already an economically competitive technology in several


niche applications. Small applications are often supplied with small batteries or
button cells. Compared to household electricity prices of around 20 cents/kWh
in Germany, for example, the costs with photovoltaics can quickly explode to
several hundred euros per kWh. It takes about 280 mignon cells to store one kilowatt
hour using high-quality alkaline manganese batteries. Now, no one would ever
think about buying 280 mignon cells to run a washing machine just once. However,
with small applications we often tend to be willing to pay whatever it costs to
buy batteries. It is often the infrequent use of these small applications that even
makes using electricity affordable. Photovoltaics can compete with this type of
high energy cost even under the cloudiest conditions. It is often an economic alter-
native even to large battery systems. However, photovoltaics will have to relinquish
its niche role if it is to become effective in protecting the climate. This will only
happen if it becomes a grid-connected system and replaces conventional power
plants.

5.5.1 What Does It Cost?


A sensible minimum quantity of power for a grid-connected photovoltaic system
is 1 kilowatt peak (kWp). There is no upper limit; this depends only on the area
available and the amount of money available. Fortunately, the prices of photovoltaic
systems are dropping so much that they are out of date even before they are
printed. This makes it difficult for any book to provide a current guideline to prices.
In 2008 it took an investment of around 4000 euros for a completely installed 1-kWp
system. The price of the photovoltaic module itself only constitutes a part of
the total system price. A good 60% of the investment goes towards the PV modules,
while the rest is spent on inverters, mounting materials, the actual mounting and
planning.
Whereas the lion’s share of the cost with photovoltaic systems is in the installation,
the payoff then comes through the electricity that is generated. The reimbursement
is usually per kWh. The operating costs of photovoltaic systems are comparatively
low. An estimated 2 to 3% of the investment costs each year will be spent on running
costs, such as insurance, possible leasing, meter rental and reserves for repairs. The
modules can normally be used for 20 to 30 years. Inverters, on the other hand,
usually wear out earlier. Repairs and replacement parts should be included in the
calculations at the outset.
Another factor that has a major effect on production costs is the return. Very
few sensible people will invest their own capital in a photovoltaic system in the
hope of at best receiving their invested capital back again over the lifetime of
the system. An investment should at least be comparable with a savings account
to make it attractive to large numbers of people. Even then a certain amount of
idealism is required, because the risks associated with photovoltaic systems are

109
5 Photovoltaics – Energy from Sand

considerably higher than those with savings accounts. For example, a flash of
lightning or a storm can completely destroy a system. If the insurance does not
pay for the damage, the investment is then lost. Another thing that can happen is
that the output is lower than projected. There may be various reasons for this. The
PV system could be the victim of poor planning, over the years a tree could
grow so that it shuts out the sun, a flock of birds could regularly be using the module
for target practice or the solar radiation may be lower than during the previous
20 years due to fluctuations in the climate. Ultimately, it is the operator of the
photovoltaic system who is responsible for dealing with all these risks. A return
of at least 6% is therefore quite appropriate for an investment not driven by
idealism.
Figure 5.17 shows the resulting production costs of a photovoltaic system based on
net investment cost. The assumption made in the calculations is that 2% of the
investment cost is for operating and maintenance costs each year and the economic
lifetime of the system is 20 years. The different coloured lines represent the calcula-
tions for various specific yields. In Germany these are normally under 1000 kilowatt
hours per kilowatt peak (kWh/kWp). For a return of 6% the production costs are
around 38 cents per kilowatt hour with a net investment cost of 3500 euros per kilo-
watt peak with a specific output of 1000 kWh/kWp. An idealist might be satisfied
with 25 cents per kilowatt hour and no return. If the output is 1500 kWh/kWp at a
location in California, at 25 cents per kilowatt hour a return of even 6% can be
achieved.

1.1 700 700


1.1
750 750
1 800 800 1
850 850
0.9 900 900 0.9
950
electricity costs in €/kWh _

950
electricity costs in € /kWh _

1000 1000
0.8 1100 1100 0.8
1200 1200
0.7 1500 1500 0.7
2000 2000
0.6 kWh/kWp kWh/kWp 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1
return 0 % return 6 %
0 0
1000 2000 3000 4000 5000 6000 7000 8000 1000 2000 3000 4000 5000 6000 7000 8000
net investment costs in € net investment costs in €/kWp

Figure 5.17 Electricity production costs based on net investment costs and specific yield for a return of
0% (left) and 6% (right).

110
5.5 Economics

5.5.2 Incentive Schemes


Whereas small-scale photovoltaic island systems are already competitive today, the
energy production costs for grid-connected photovoltaic systems are in most cases
still higher than normal market prices. It currently only makes sense to install large
numbers of grid-connected photovoltaic systems if state incentive schemes are avail-
able. In Germany, compensation is regulated through the Renewable Energy Law
(Erneuerbare Energien Gesetz, EEG). Other countries have also adopted the com-
pensation principle. A fixed price is stipulated for each kilowatt hour a photovoltaic
system feeds into the public electricity grid. The relevant electricity company pays
this fee but is allowed to split any associated additional costs among all its electricity
customers. The law is aimed at making solar power competitive. The law therefore
provides an annual degression for systems in buildings. The compensation for new
systems thus drops each year (Table 5.5). The increased compensation is guaranteed
for 20 years after a system has been installed. After this, the fee drops to normal
market prices.

Table 5.5 Compensation in Germany for grid-connected photovoltaic systems in cents/kWh


based on the Renewable Energy Law.

Commissioning 2005 2006 2007 2008 2009 2010a 2011a 2012a


Building systems 54.53 51.80 49.21 46.75 43.01 39.57 36.01 32.77
As of 30 kW 51.87 49.28 46.82 44.48 40.91 37.64 34.25 31.17
As of 100 kW 51.30 48.74 46.30 43.99 39.58 35.62 32.42 29.50
As of 1000 kW 51.30 48.74 46.30 43.99 33.00 29.70 27.03 24.59
Façade bonus 5.00 5.00 5.00 5.00 0.00 0.00 0.00 0.00
Open space systems 43.42 40.60 37.96 35.49 31.94 28.75 26.16 23.81
a
Compensation rates for 2010 to 2012 can vary slightly depending on total installed power for the previous year.

The EEG law has succeeded in triggering a photovoltaic boom in Germany since
2004. Other countries, including Spain, have adopted the same type of law. There
are countries that use other incentive models, such as prescribed quotas for renew-
able energies. However, compared to the German EEG, these models have not
proven to be very successful.
If, as predicted, the costs go down, photovoltaics at certain locations could become
economically competitive in less than ten years without an increase in compensa-
tion. Many countries that have high solar radiation levels experience a large increase
in electricity demand because of air-conditioning systems used in the summer. Due
to the associated high electricity costs in summer, photovoltaics can become
economically viable even earlier in these countries. In many countries the high
initial investment required to install photovoltaic systems is made easier because
low-interest loans or grants are available. However, the conditions and programmes
for these are often subject to change.

111
5 Photovoltaics – Energy from Sand

From the Quotation for a PV System to the Running Installation

■ Find out about public grants and low-interest loans.


■ Review economic feasibility based on quotations received (see Figure 5.17).
■ Clarify connection conditions with the energy supply company.
■ Arrange for the system to be installed by company specializing in photovoltaics.

5.6 Ecology

There is a persistent rumour that it takes more energy to produce the silicon for a
solar cell than the cell could ever generate during its lifetime. No one knows where
this rumour started. Presumably it was spread by opponents of solar technology
before detailed studies on the subject were even available.
It is true that quite a lot of energy is needed to produce solar cells. Temperatures
of well over 1000 °C are required in the production of silicon and for its subsequent
purification. However, in contrast to conventional coal-fired, natural gas and nuclear
power plants, no other energy supplies are needed in the operation of photovoltaic
systems. Once a solar system is in operation, it begins to supply back the energy
used in its production. Various scientific studies in the 1990s showed that in Germany
it takes around three to four years for solar cells to regenerate the energy needed to
produce them (Quaschning, 2005), but less than two years in Southern Europe or
at good US sites. Due to increased efficiency and decreases in cell thickness, these
figures will now be much lower. With a lifetime of 20 to 30 years, a solar cell
generates vastly more energy than that used to produce it.
Today a 0.3 mm-thick, 6-inch crystalline silicon solar cell weighs around 16 grams.
The amount of energy needed to produce solar cells will continue to fall, because
there is an active interest in drastically reducing the materials used, and, conse-
quently, in reducing cost. For example, the amount of material used in thin film
cells has already decreased considerably. It is expected that in the future solar
systems will be able to make up the energy needed for their production in just a
few months.
Various chemicals, some of them toxic, are used to produce solar cells. As with all
chemical systems, care has to be taken that no chemicals escape into the environ-
ment. This should not be a problem with the modern production systems and high
environmental standards that exist today.
An old solar system that is no longer useable, on the other hand, is considerably
less problematic. Nevertheless it would be a shame simply to scrap old and unviable
solar systems, because they contain valuable raw materials. The solar industry is
therefore working on methods to recover these materials. Initial attempts at recy-
cling solar cells seem promising. The solar modules are dismantled into their origi-
nal parts, and new photovoltaic modules are produced from the reclaimed cells.

112
5.7 Photovoltaic Markets

5.7 Photovoltaic Markets

Among the renewable energies that currently exist, photovoltaics offers the most
possibilities for different uses. The advantage is the modular structure of these
systems. Almost any desired generator size is possible, from the milliwatt range for
supplying power to pocket calculators and watches all the way to the megawatt
range for supplying public electricity. Pocket calculators supplied by photovoltaics
were already in widespread use decades ago.
Whereas in 1980 85% of all solar modules were being produced in the USA, this
share shrank to less than 10% by 2005. Japan, China and Germany have replaced
America as the leaders in photovoltaics.
Large-scale systems that feed solar power into the public grid first became popular
after successful marketing programmes in Japan and Germany. Government schemes
in both countries boosted the production of photovoltaic modules and have contrib-
uted towards annual market growth rates between 20 and 40% since the early 1990s
(Figure 5.18). In the meantime, other countries like Spain, Italy and the USA have
also created attractive conditions for the installation of grid-connected photovoltaic
systems.

16000 MW
Germany
14000 Japan
USA
Spain
12000
Others
10000

8000

6000

4000

2000

0
1992

1993

1994

1995

1996

1997

1998

1999

2000

2001

2002

2003

2004

2005

2006

2007

2008

Figure 5.18 Development of overall photovoltaic power installed worldwide.

In Germany, the popularity of grid-connected photovoltaic systems did not begin


until the early 1990s with the so-called 1000-roof programme. With the help of state
schemes more than 2250 photovoltaic systems were erected, mainly by private
households. When the programme was phased out, the use of photovoltaics began
to stagnate. It was not until 2000 that fresh impetus was given through the introduc-
tion of the Renewable Energy Law.
With over 500 000 photovoltaic systems and an output of more than 5300 megawatts,
Germany in 2008 had the most systems in operation by a considerable margin. These

113
5 Photovoltaics – Energy from Sand

systems fed 4.3 billion kilowatt hours of electric power into the grid, which was a
0.7% share of the overall electricity supply. Photovoltaics was thereby supplying
more than one million households with carbon-free electricity. Photovoltaics com-
panies in Germany generated a turnover of 7 billion euros and used it to create 48 000
future-orientated workplaces. The Federal Association for Solar Economy is counting
on more than 100 000 workplaces by 2020 (Bundesverband Solarwirtschaft, 2009).

5.8 Outlook and Development Potential

Even if the quantity of solar power is still relatively small, in the medium term
photovoltaics will be able to provide the largest share of environmentally compatible
electricity supply. From a purely mathematical standpoint, it could be used to supply
the world’s entire energy needs. This would only take a fraction of the surface of
the Sahara Desert to accomplish. Even countries like Great Britain, Germany and
France would be able to cover all their electricity requirements through photovolta-
ics. On the other hand, from a technical perspective it is not a good idea to rely
solely on one technology for the future supply of energy. Photovoltaic systems work
well in combination with other renewable energy systems, such as wind power,
hydropower and biomass systems. A well-planned combination of systems will
increase supply reliability and avoid the building of large storage systems to ensure
sufficient supplies are available at night and during the winter.
Costs will have to drop further before large numbers of countries start using pho-
tovoltaics on a considerably larger scale than at present. Past experience has shown
that major cost reductions are possible. Whereas the price of photovoltaic modules
was still around 60 inflation-adjusted US dollars per watt in 1976, by 2007 it had
already dropped to around 3 dollars per watt (Figure 5.19).

100
1976
$2000
Wp
photovoltaic module prices

1980
1990
10

2007
2000

2020
1

0.1
0.0001 0.001 0.01 0.1 1 10 100 GWp
cumulated photovoltaic module production

Figure 5.19 Development of inflation-adjusted photovoltaic module prices on the basis


of total quantity of modules produced worldwide.

114
5.8 Outlook and Development Potential

What is crucial for cutting costs is an increase in production. If production quantities


rise, then costs will drop noticeably due to the effects of streamlined production and
also because of technical advances. During the past 30 years cost savings of around
20% have been achieved due to a doubling in the total quantities of photovoltaic
modules produced. There is nothing to indicate that this development will not con-
tinue. It is possible that the prices of photovoltaic modules could fall below US$1
per watt by 2020. As a result, the current cost to generate electricity using photo-
voltaic systems would have shrunk to about one fourth.
This would then make photovoltaic systems very interesting to end users in Central
Europe, even without the need for any government subsidy schemes. A photovoltaic
system would be able to produce electricity more cost-effectively for household use
than an energy supply company could deliver to the household. In the sunniest parts
of the world photovoltaics could produce energy more cost-effectively than any
conventional alternatives. For that reason, the main markets for solar energy in the
long term will be in places other than Western Europe.

115
6
6 Solar Thermal Systems –
Year-Round Heating from
the Sun

The light and warmth of the sun give us a special sense of well-being, which is why
summer is eagerly awaited in cool parts of the world such as Northern Europe. Most
people in these countries are also lucky enough to be able to reproduce these condi-
tions even in the depths of winter. Our homes and workplaces are heated to a
comfortable level and well lit, and our water is heated so that we can enjoy hot
baths and showers at any time. The luxury of always being able to set the desired
temperature is taken for granted today, one of the most pleasant achievements of
our prosperous society. It is difficult to imagine the period after the Second World
War, when not enough fuel was available in the cold winters to maintain even
reasonably bearable temperatures in our dwellings. However, our prosperity and
fossil fuels have eliminated these conditions once and for all – not everywhere in
the world but at least for us.
But even eliminating fossil fuels will in no way jeopardize our privilege of always
being able to select the temperature we want. Solar thermal energy – the heat from
the sun – is an important alternative. When sunlight shines through the windows,
the power of the sun is already helping to warm up the building. For centuries the
sun has been a major source of heat for our homes.
In 1891 the metal manufacturer Clarence M. Kemp from Baltimore was awarded
the first patent in the world for a technical solar thermal system. This was a very
simply constructed storage collector for heating water. In 1909 the Californian
William J. Bailey introduced an optimized system concept that separated the solar
heat collector from the water storage cylinder. Solar heating systems were marketed
successfully in certain regions until the Second World War, after which the market
collapsed because of competition from fossil fuel.

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

116
6.1 Structure and Functionality

Figure 6.1 Modern solar thermal collector systems are an important alternative to conventional oil and
natural gas heating. Photos: www.wagner-solar.com.

It was not until the oil crisis in the 1970s that solar thermal power was rediscovered.
In the years that followed there were still many teething problems, and not all
systems ran smoothly. Today a variety of solar thermal system variants are avail-
able, and these systems are much more sophisticated than in the past (Figure 6.1).
In combination with optimized heat insulation and other renewable heating systems,
such as biomass heating and renewably operated heat pumps, these systems can
make an important contribution towards the supply of carbon-free heat.

Solar Collectors, Solar Absorbers, Solar Cells or Solar Modules


– What Next?
These terms are often used interchangeably, so to avoid confusion it’s impor-
tant to be clear about what exactly they refer to.
A sun collector or solar collector is used to extract heat from the radiation of the sun.
Thus, a collector always makes something hot. At the heart of a solar collector is a solar
absorber. The solar absorber absorbs the radiation of the sun and converts it into heat.
Solar collectors are used to heat domestic water supply, supplement room heating and
produce high-temperature process heat. Thermal power plants can even generate electric-
ity from high-temperature heat (see Chapter 7). However, even then solar collectors are
first used to produce the heat.
A solar cell is a photovoltaic cell that converts solar radiation directly into electric energy
(see Chapter 5). Although a solar cell can also get hot, it is different from a solar col-
lector in the sense that this is an undesired side effect. The heat actually reduces effi-
ciency during the production of electricity. A solar module consists of a large number
of solar cells and also generates electric energy from sunlight.

117
6 Solar Thermal Systems – Year-Round Heating from the Sun

6.1 Structure and Functionality

Solar thermal energy is a technology for converting solar radiation into heat. The
field of application for this technology is extensive. The higher the temperatures are
required to be, the more complicated the technical implementation. The principle is
similar with all solar thermal systems. A solar collector first collects the sunlight.
The term collector comes from the Latin word ‘collegere’, which means to collect.
The main component of a collector is the solar absorber. It absorbs the sunlight and
converts it into heat (Figure 6.2). It then transfers this heat to a heat transfer medium.

Figure 6.2 Processes in a solar flat-plate collector.

The heat transfer medium may simply be something like water, air or even oil or
salt. Heat loss is unavoidable during the conversion. Part of the solar radiation is
reflected and does not even reach the absorber. A certain amount of heat is lost
before it can be transferred to the heat transfer medium. The trick is to construct a
collector so that heat loss is as low as possible but the collector is still cost-effective
to produce. Appropriate collectors should be used depending on the field of applica-
tion and desired temperatures.
The efficiency curve describes the performance of a collector. Research institutes
measure collectors under strictly defined conditions to determine the curve. These
characteristic curves can then be obtained from the manufacturer of the collector or
over the Internet.

■ http://www.spf.ch/spf.php?lang=en SPF Collector Test Reports


■ http://www.itw.uni-stuttgart.de ITW Kollektor-Testberichte (in German)

118
6.1 Structure and Functionality

Collector Efficiency

The efficiency and efficiency curve of a solar collector can be determined using the
three parameters η0, k1 and k2. The optical efficiency η0 describes which portion of
sunlight is converted into heat by the absorber. The absorber itself or the front panel
of glass actually reflects a portion of the sunlight before it can even be absorbed.
Depending on the type of collector, optical efficiency is between 70 and 90%. The
two loss coefficients k1 and k2 indicate how high the heat loss is in the collector.
The hotter the collector is, the higher the heat loss and the less useful heat the col-
lector can emit. High loss coefficients also mean high heat loss. The formula for
collector efficiency is as follows:
k1 ⋅ Δϑ + k2 ⋅ Δϑ 2
η = η0 − .
E

Here Δϑ indicates the temperature difference between the collector and the environ-
ment and E the intensity of the solar radiation.
With an ambient temperature of 25 °C and a collector temperature of 55 °C, the
result is a temperature difference of Δϑ = 30 °C, or 30 K. On a pleasant summer’s
day with a solar radiation intensity of E = 800 W/m2, the calculation of a flat-plate
collector with an optical efficiency of η0 = 0.8 and the loss coefficients k1 = 3.97 W/
(m2 K) and k2 = 0.01 W/(m2 K2) gives a collector efficiency of
2 m K 2 ⋅ (30 K )
3.97 mW2K ⋅ 30 K + 0.01 W 2

η = 0.8 − = 0.64 = 64%.


800 mW2

A collector 4.88 m2 in size then supplies 2500 watts of power. This is sufficient to
heat 100 litres of water from 33.5 °C to 55 °C in one hour.

The collector efficiency curve describes the progression of efficiency based on the
temperature difference between the collector and the environment (Figure 6.3). It
shows that efficiency drops as the temperature rises until the collector reaches an
efficiency of zero and finally cannot supply any more power.
Almost all thermal solar systems require storage in addition to a collector. There
are very few cases when the sun shines at exactly the same time as heat is needed.
Even a simple water tank can function as a thermal store and should be well insu-
lated to reduce heat loss. The storage size depends mainly on the heat requirement
and on how long the heat needs to be stored. Daytime storage cylinders for hot water
systems in single-family homes are designed to bridge a few days of demand and
usually only hold a few hundred litres. If, in addition to hot water, very large quanti-
ties of heat for heating are also to be stored, then seasonal heat storage cylinders
are required. These store heat in the summer and then release it again in the winter.
Large heat storage cylinders for small housing estates reach sizes of several hundred
or even thousand cubic metres. Large storage cylinders generally have less specific

119
6 Solar Thermal Systems – Year-Round Heating from the Sun

100 %

90 % optical losses

collector efficiency 80 %

70 %
thermal losses
60 %

50 %

40 %

30 %

20 %

10 %

0%
0 10 20 30 40 50 60 70 80 90 100 110 120 130
temperature difference collector - ambient air in °C

Figure 6.3 Collector efficiency curve.

heat loss than small ones. The storage volume increases considerably faster than the
size of the storage surface. However, storage losses depend only on the size of the
surface of the storage.

6.2 Solar Collectors – Collecting the Sun

6.2.1 Swimming Pool Absorbers


Different types of collectors are available for different purposes. The simplest type
of collector consists of only one absorber. Placing a dark garden hose filled with
water in the sun for a time produces enough heated water for a short but warm
shower. A garden hose thus already has the characteristics of an absorber. In prin-
ciple, a professional swimming pool absorber is also nothing more than a simple
black plastic pipe that absorbs the sunlight almost optimally due to its dark colour.
Weather-resistant plastics that cope well with UV light and chlorinated pool water
are used for this purpose. The materials that are suitable are polyethylene (PE),
polypropylene (PP) and ethylene-propylene-dien-monomere (EPDM). PVC should
be avoided for ecological reasons. In winter it is futile trying to use a garden hose
to extract warm shower water. The same applies to swimming pool absorbers.
Although an absorber continues to absorb the sun’s radiation, the heat loss in the
absorber pipe itself is so great that hardly any heat can be drawn from the end of a
pipe. Technically more sophisticated collectors are required to use the sun’s heat
during winter and the transitional periods of the year or when water temperatures
should be warmer than the water in a swimming pool.

120
6.2 Solar Collectors – Collecting the Sun

6.2.2 Flat-Plate Collectors


With flat-plate collectors a front panel of glass reduces heat loss considerably.
Unfortunately, the front glass also reflects some of the sunlight. Therefore, when
collector temperatures are very low, the efficiency of a swimming pool absorber can
be higher than that of a flat-plate collector. However, flat-plate collector efficiency
increases considerably as temperatures rise.
In summer there are times when solar thermal systems do not need any heat. For
example, when a heat storage cylinder is completely full the system does not pump
any more water through the collector. This is referred to as a collector standstill.
The temperatures in collectors can easily rise above 100 °C. Very good flat-plate
collectors reach standstill temperatures between 150 °C and 200 °C. This therefore
rules out the use of plastic pipes for absorbers. The absorbers of flat-plate collectors
usually consist of copper pipes fixed to a thin metal plate (Figure 6.4). The absorber
itself is located in a collector enclosure that is well insulated at the back to minimize
rear heat loss.

Figure 6.4 Cross-section of a flat-plate collector. Picture: BBT Thermotechnik GmbH.

Metallic materials by nature do not have a black surface that absorbs the sun’s
radiation well, so they need to be coated. The first option that is usually considered
for the coating is black paint. Although temperature-resistant black paint is good
for this purpose, there are other far better materials for absorber coating. When a
black surface gets warm, it sends out some of the heat energy in the form of thermal
radiation. This can be observed with an electric hot plate that has been switched on.
One can feel the heat radiating from the hot plate without even touching it. The
same effect occurs with a black-painted absorber. The absorber only radiates part
of its heat to the water flowing through. Another part is emitted back into the envi-
ronment as undesirable thermal radiation.
Thermal radiation losses can be minimized through the use of selective coatings
(Figure 6.5). These coatings absorb the sunlight just as well as a black-painted plate.

121
6 Solar Thermal Systems – Year-Round Heating from the Sun

However, they emit a much lower amount of thermal radiation. Materials for selec-
tive coatings can no longer simply be painted or sprayed onto a surface. More
complicated coating methods are now necessary.

Figure 6.5 Principle of selective absorbers.

6.2.3 Air-Based Collectors


In most cases, solar collectors heat water. However, air and not water should be
used to heat the air of rooms. With conventional heating systems, radiators or
heating pipes installed under the floor transfer the heat produced from hot water to
the air. Instead of water, air can also be transmitted by a solar collector.
As air absorbs heat considerably less effectively than water, much larger absorber
cross-sections are required. Otherwise, in principle, air-based collectors differ very
little from liquid-based flat-plate collectors. Figure 6.6 illustrates an air-based col-
lector with perforated absorbers. An integrated photovoltaic module can deliver the

solar safety glass

photovoltaic cells
absorber ribs

insulation intake opening

outlet tube

Figure 6.6 Cross-section of an air-based collector. Illustration: Grammer Solar GmbH.

122
6.2 Solar Collectors – Collecting the Sun

electricity to drive the fan motor. Air-based collectors are a particularly interesting
alternative as providers of support heating. However, heat storage is more difficult
with systems that use air-based collectors.

6.2.4 Vacuum-Tube Collectors


With flat-plate collectors the air between the absorber and the front glass panel is
the source of most of the heat loss. It causes convective heat loss and continuously
transports the heat from the absorber to the glass panel. This then emits the heat
unused back into the environment.

How a Blown-Up Plastic Bag Evacuates Its Air


The term vacuum originates from the Latin word ‘vacus’, which means
‘empty’ or ‘free’. A vacuum is generally thought of as space empty of matter.
However, it is practically impossible to produce a totally air-less space on earth.
In technology and in physics a vacuum is interpreted merely as air pressure that is con-
siderably lower than normal air pressure. To produce a vacuum, one uses a vacuum
pump to pump the air out of a stable space. In principle, it is also possible to use one’s
mouth to produce a rough vacuum – for example, by sucking the air out of a glass bottle.
However, a plastic bag cannot be evacuated – at least not in any normal environment
on earth. On the other hand, an almost perfect vacuum exists in space. If one opens an
empty plastic bag in space and then closes it again, there will also be a vacuum in the
blown-up bag. If one returns to earth with the bag, the ambient air pressure compresses
it again.
Ambient air pressure originates through the power of the weight of air columns above
the earth’s surface. The atmospheric air weighs around 10 tons per square metre of earth
surface. The question is, why doesn’t this enormous pressure simply smash the glass of
a collector? The answer is simple: the space under the glass is also filled with air, which
produces the necessary counterpressure. On the other hand, if the air in the space behind
the glass is evacuated, the glass will normally bend and shatter.

Heat losses can be reduced considerably if a vacuum exists between the absorber
and the front glass because of the air movement in the collector. This is the principle
used by vacuum flat-plate collectors. As the outer air pressure would press the front
covering against the absorber, spacers are needed between the underside of the col-
lector and the glass covering. The vacuum cannot be kept stable for a long period
because penetration of air where the glass and collector housing intersect cannot be
completely prevented. Therefore, vacuum flat-plate connectors must be evacuated
again at certain intervals. This is done by attaching a vacuum pump to a special
valve on the collector. These drawbacks were the reason why vacuum flat-plate
collectors never became popular.
Another type of collector available is the vacuum-tube collector, which does not
have the same disadvantages as the vacuum flat-plate collector. With vacuum-tube
collectors, a high level of vacuum is completely enclosed in a glass tube.

123
6 Solar Thermal Systems – Year-Round Heating from the Sun

Consequently, these collectors are considerably easier to produce and maintain for
long periods than vacuum flat-plate collectors. The shape of the glass tubes enables
them to withstand outside air pressure better so that metal rods are not necessary
for support.
In the enclosed glass tube of a vacuum tube collector is a flag absorber, which is a
flat metal absorber that has a heat pipe integrated in the middle (Figure 6.7 left). A
lightly vaporizing medium like methanol is enclosed in this heat pipe. If it vaporizes
due to the heat of the sun, the vapour rises upwards. The heat pipe sticks out of the
glass tube at the top end. This is where the condenser, which condenses the heat
medium again, is located. In the process, it transfers the heat energy over a heat
exchanger to the water flowing through. After condensation the once-more liquid
medium in the heat pipe flows downwards. For maximum functionality the tubes
should be mounted with a certain minimum tilt.

Figure 6.7 Vacuum-tube collectors. Left: Collector with heat pipe. Right: Tubes with direct flow.
Illustrations: Viessmann Werke.

Vacuum-tube collectors with passing heat carrier pipes (Figure 6.7 right) are also
available. With this system the heat carrier liquid flows directly through the collec-
tor. A heat exchanger is then not required and the collector does not necessarily
have to be mounted at an angle.
Because hydrogen molecules are extremely small, atmospheric hydrogen cannot be
totally prevented from penetrating the vacuum. With time this also destroys the
vacuum. So-called getters, which can chemically bind the hydrogen that penetrates
the vacuum over a long period of time, are built into a collector to prevent this from
happening.
The advantage of vacuum-tube collectors is a very high energy yield, especially
during the cooler seasons of the year. Compared to one with normal flat-plate
collectors, a solar system with vacuum-tube collectors requires less space for
the collector. The disadvantage is the substantially higher cost of the collector
(Figure 6.8).

124
6.3 Solar Thermal Systems

Figure 6.8 Comparison of flat-plate and vacuum-tube collectors. Illustration: Viessmann


Werke.

6.3 Solar Thermal Systems

6.3.1 Hot Water from the Sun

6.3.1.1 Gravity Systems


The sun can be used to heat water to a high temperature. This is something we
learned from Archimedes, who used a convex mirror to bring water to the boil as
far back as 214 BC. That’s not surprising, some will say, because Archimedes was
not from Hamburg, London or Vancouver. In Southern Italy it is simply easier to
use the sun’s energy. This is quite correct, but only up to a point.
The temperatures in the Mediterranean region are much higher than in Germany,
Great Britain or Canada and there is no worry there about frost damage to hot water
systems. Systems that use solar energy to produce hot water can be constructed
more simply and thus more cheaply in frost-free regions than in colder climates.
However, even at high latitudes solar energy is an excellent option for producing
desired heat.
Unfortunately, not all the people in sunny countries have received the message that
they should be making use of solar energy. Thermal solar systems are common in
Cyprus and Israel but not in Spain or the southern USA, although these areas also
have a sunny climate. Gravity systems are mainly used in countries that do not
get frost (Figures 6.9 and 6.10). A flat-plate or vacuum-tube collector collects the
solar radiation and heats water that flows through the collector. A hot water storage

125
6 Solar Thermal Systems – Year-Round Heating from the Sun

cylinder is installed at a higher level than the collector. As cold water is heavier
than warm water, it sinks down from the storage cylinder to the collector. Here the
water is heated by the sun and then rises to the top until it reaches the storage cyl-
inder again. If no more solar radiation is available, the cycle stops until the sun
starts it up again.

Figure 6.9 Left: Demonstration model of a gravity system. Right: Gravity system in Spain.

Figure 6.10 Solar gravity system (thermo-syphon system).

If properly designed, this kind of solar system can cover almost all the hot water
needs of a family living in a warm, sunny southern climate. It is only during some
of the sun-starved weeks of the year that the water will not completely reach desired
temperatures. Either this is accepted in those countries or a supplemental heater, for

126
6.3 Solar Thermal Systems

example, with natural gas, is installed. If on occasion the heat of the sun is not suf-
ficient, a supplemental heater then takes over.
In China solar thermal systems for heating water have gained a high market share.
In remote rural regions there is almost no access to fossil fuels. The inhabitants of
these regions do not have the possibility of heating their water as and when required
if there is no sun. China uses mainly vacuum-tube collectors because of the need
to guarantee high supply reliability from solar systems.

6.3.1.2 Systems with Forced Circulation


A system should be technically optimized if the sun is to be used to produce hot
water in areas with low outdoor temperatures. It would be too risky to heat up water
for domestic use directly in a collector because the water would freeze in winter
and therefore could destroy the collector. Consequently, the water that flows through
the collector is mixed with an antifreezing agent. However, antifreezing agents
cannot be used in the water supply as they have negative effects on health. A heat
exchanger therefore separates the water circulation from the solar circulation and
transfers the heat to a hot water storage cylinder. The storage cylinder is normally
designed so that it can provide enough hot water for two to three bad days until a
major supply of heat comes from the solar collector.
Flat roofs are less common in Central and Northern Europe and North America than
they are in Southern Europe (Figure 6.11). Hot water storage cylinders are tradition-
ally located in a basement or in a utility room. If the hot water storage is situated

Figure 6.11 Single-family house with photovoltaic system (left) and flat-plate collectors
for heating water (right). Photo: SunTechnics.

127
6 Solar Thermal Systems – Year-Round Heating from the Sun

below the collector, the water has to be pumped through the collector. A pump moves
the water in a solar circulation through the collector and a control mechanism ensures
that the pump only starts when the temperature of the collector is higher than that
of the storage cylinder (Figure 6.12). A conventional boiler heats up the water during
the transitional seasons and in winter so that a hot shower is possible all year round.
It can happen that during the summer a solar collector will heat up an entire storage
cylinder to a predetermined maximum temperature. This maximum temperature is
usually set at 60 °C to prevent large deposits of lime from forming. If the cylinder
is full, the control interrupts the incoming supply from the collector. Despite full
irradiation from the sun, no more water will flow through the collector. The collector
can then heat up to temperatures well over 100 °C and evaporate the water. If the
expansion tank is large enough, it absorbs the expansion of the water volume.

Figure 6.12 Pumped solar thermal system for heating domestic hot water.

6.3.2 Heating with the Sun


Solar thermal systems can be used to provide not only hot water but also heating.
In principle, this only requires increasing the sizes of the collector and the storage
cylinder and connecting them to the heating cycle. As no provision is normally made
for domestic hot water in the heating cycle, two separate heat storage cylinders are
needed – one for hot water and the other for drinking water. Combination boilers
can integrate both storage elements and reduce heat losses (Figure 6.13).
In places like Germany and Great Britain the sun is only able to meet heating
demand during the transitional periods of spring and autumn. Consequently, these
systems are usually designed so that solar energy only serves as a supplemental heat
supply. The output of the collectors is not sufficient to cover the heat needed during
the winter. In principle, the sun could be used to cover all heating requirements.
However, this would necessitate either an extremely large collector or an extremely

128
6.3 Solar Thermal Systems

Figure 6.13 Solar thermal system for heating up domestic hot water and providing
support heating.

large storage cylinder that stores thermal heat from the summer for the winter. The
cost of the solar system would increase considerably as a result. Economically, it
currently makes more sense to integrate only one storage cylinder with a few days’
capacity into a system. This would enable between 20% and 70% of the heat require-
ment to be covered by the sun (Figure 6.14).

Figure 6.14 Large roof-integrated solar thermal system for heating water and providing
supplementary heating. Photo: SunTechnics.

129
6 Solar Thermal Systems – Year-Round Heating from the Sun

6.3.3 Solar Communities


If many houses in a community have solar collectors, these can be integrated into
a solar district heat network. This can also involve setting up a large central collector
complex. At the heart of such a heat network is a central heat storage tank. Its size
helps to minimize heat losses, thus also enabling heat to be stored for longer periods.
However, the extensive tube systems involved can result in disadvantages, such as
higher costs and the possibility of major line losses. Some solar district heat com-
munities have already been successfully installed (Figure 6.15).

Figure 6.15 Solar district heat supply.

6.3.4 Cooling with the Sun


As paradoxical as it may sound, the heat from the sun can also be used to provide
excellent cooling for buildings. In the hot and sunny regions of the world large
numbers of energy-hungry air conditioning systems ensure that room temperatures
are pleasantly cool. The sunnier and hotter it is, the greater the need for cooling. As
the radiation from the sun increases, the output of a thermal collector also increases.
In contrast with the requirement for heat, cooling load demand coincides almost
perfectly with the supply of the sun.
Along with a large and efficient collector, an absorption-refrigerating machine is an
essential element of a solar cooling system (Figure 6.16). In this context, the term
absorption does not refer to a solar absorber. An absorption-refrigerating machine
utilizes the chemical process of sorption. A chemist interprets sorption or absorption
as the absorption of a gas or a fluid by another fluid. A popular example is the dis-
solution of carbon dioxide gas in mineral water.
Absorption-refrigerating machines use a sorbable cooling agent with a low boiling
point, such as ammonia, which is later dissolved in water. Even aside from ammonia,

130
6.3 Solar Thermal Systems

Figure 6.16 Principle of solar cooling with absorption-refrigerating machines.

water itself under low pressure can be used as a cooling agent. Lithium-bromide is
then suitable as the solvent.
The cooling agent boils in a vaporizer at low temperatures. In the process it extracts
the heat from the cooling system. The cooling agent then has to be liquefied again
so that it can provide continuous cooling through renewed evaporation. With the
help of a few tricks and an indirect way through sorption, the liquidation also suc-
ceeds through the use of solar heat.
The cooling machine absorber first mixes the cooling agent vapour with the solvent.
This produces sorption that releases the heat. This heat is either used to heat the
water or is transferred to a cooling tower.
A solvent pump transports the liquid solution, which has been enriched with a
cooling agent, to the generator. The generator separates the cooling agent and the
solvent on the basis of their different boiling points. Heat from efficient solar col-
lectors is used for the boiling process. Temperatures of 100 °C to 150 °C are optimal.
The separated vapour-formed cooling agent then enters a condenser, which liquefies
the cooling agent. The condensation heat is also dissipated either as useful heat or
over a cooling tower. The liquid cooling agent enters the vaporizer over an expan-
sion valve and the solvent returns again to the refrigerating machine absorber. The
cooling agent cools down considerably through the expansion in the expansion valve
and can again deliver cooling to the cooling system over the vaporizer.

6.3.5 Swimming with the Sun


During the outdoor swimming season in Central Europe water temperatures normally
reach 10 to 19 °C. The water is only warmer than this for a few days during the height
of summer, although this period is becoming longer because of global warming.
These water temperatures are a result of swimming pool water being heated by the
sun. The following example shows how enormous the energy content of the sun is.
The example uses Lake Constance, a popular lake for swimming that is visited by

131
6 Solar Thermal Systems – Year-Round Heating from the Sun

millions of tourists annually. In winter the lake cools off quite considerably and in
1880 and 1963 it was even completely frozen over. In summer, in contrast, it reaches
a temperature high enough to attract hordes of swimmers. If the entire supply of coal
consumed in Germany each year were used to heat the water, it would only be enough
to heat up the 50 cubic kilometres of water of Lake Constance just once to a tem-
perature of 9 °C. Looking at the USA, the whole US American primary energy supply
of one year would only suffice to heat up Lake Michigan by less than 5 °C on one
occasion. The sun, on the other hand, can provide pleasant water temperatures
without a problem and do it over many weeks.
Despite the heat of the sun, many bathers still find swimming pool temperatures too
low in summer, so that pools are artificially heated. And because we only have to
heat up a small swimming pool and not all of Lake Constance or Lake Michigan,
many places pull out all the stops. In Germany there are about 8000 public outdoor
and indoor swimming pools. In addition, there are around 500 000 private pools. In
the USA there are about 8.6 million pools in total and 270 000 commercial pools.
Several hundred million euros are spent each year alone on heating public pools.
Yet alternatives are available that would definitely enable us to use the sun to save
fuel costs and considerably reduce the emission of carbon dioxide.
Outdoor swimming pools are particularly suitable targets for the use of solar energy.
Simple swimming pool absorbers heat up pool water directly. A pump conveys the
water through an absorber and a simple control ensures that water is only pumped
when the sun actually can heat it up (Figure 6.17). If the water were pumped through
the absorber tube at night as well, this would result in the pool water being cooled
again. A good half-square metre of solar absorber surface is needed per square metre
of pool area. The surface space that is needed is often found on buildings near the
swimming pool. Covering a pool at night can save further energy.

Figure 6.17 System for swimming pool heating using solar energy.

132
6.4 Planning and Design

6.3.6 Cooking with the Sun


In many countries cooking is still often done on an open wood fire. About 2.5 billion
people around the world use this traditional method to prepare their meals. From
the energy point of view, however, an open fire is anything but efficient. Firewood
does not last long. In many countries wood is cut down at a faster rate than trees
can grow back. Furthermore, the smoke produced by open fires is responsible for
many illnesses. In sunny countries solar cookers offer an alternative to traditional
hearths.
A solar cooking box is a very simple cooking system: a wooden box painted black
inside that is covered with a sheet of glass angled in the direction of the sun. This
very simple solar collector is actually capable of heating water and food. However,
it is not very efficient and the glass cover makes it difficult to prepare food. An
efficient solar cooker is a neater solution. With a solar cooker, the cooking pot
is situated in the middle of a convex mirror (Figure 6.18). The mirror is moved
approximately every quarter-hour so that it is directed optimally towards the sun.
With a mirror diameter of 140 cm and good sun radiation, it is possible to bring
three litres of water to the boil in about half an hour.

Figure 6.18 Solar cooker in Ethiopia. Photo: EG Solar e.V., www.eg-solar.de.

6.4 Planning and Design

Of all the solar thermal systems described, the systems that use solar heat to heat
up domestic water and to supplement other heating systems are the most widely
used. The planning tips are therefore limited to these two variants.

133
6 Solar Thermal Systems – Year-Round Heating from the Sun

6.4.1 Solar Thermal Heating of Domestic Hot Water

6.4.1.1 Outline Design


In Germany, Britain and other temperate regions, solar thermal domestic hot water
systems are normally designed so that on a yearly average the sun covers 50 to 60%
of the hot water requirement. As the amount of sunshine in these regions fluctuates
considerably during the year, a solar system can usually only cover the total hot
water demand during the summer months. In the winter, the sun’s share can fall
below 10% (Figure 6.19). A conventional heating system then has to cover the rest.
In sunny regions like California and Southern Europe the solar share can easily
reach more than 80%.

100 %

90 %

80 %

70 %
Solar f raction

60 %

50 %

40 %

30 %

20 %

10 %

0%
Jan. Feb. March April May June July Aug. Sep. Oct. Nov. Dec.

Figure 6.19 Typical solar share of solar thermal domestic hot water systems in Berlin,
London and Los Angeles (5 m2 flat-plate collector in Berlin and London, 4 m2 in Los Angeles,
300 l storage cylinder, 2700 kWh hot water demand).

A simple rule of thumb for designing solar thermal systems to provide domestic hot
water, based on the number of people in a household, is:
■ Collector size: 1–1.5 m2 flat-plate collectors per person
■ Storage size: 80–100 litres per person
If vacuum-tube collectors are used, the collector size can be around 30% smaller.
The collectors should not be less than three to four square metres in area because
below this size the losses in the tubes increase to above average.

6.4.1.2 Detailed Design


Hot water requirements have to be determined before a detailed design can be drawn
up. Ideally, the system should include a hot water meter that provides a direct
reading of actual consumption or a record listing hot water use over a long period

134
6.4 Planning and Design

of time. As a rule of thumb, requirements can be approximated based on a hot water


temperature of 45 °C:
■ Low usage: 15–30 litres or 0.6–1.2 kWh per person per day
■ Medium usage: 30–60 litres or 1.2–2.4 kWh per person per day
■ High usage: 60–120 litres or 2.4–4.8 kWh per person per day

Storage Size and Hot Water Demand

The storage size Vstorage should be equal to about twice the total requirements of P
persons with a daily requirement Vperson per person:
Vstorage = 2 ⋅ P ⋅Vperson .
Based on the hot water requirement Qperson per person per day, the annual hot water
requirement QHW for the supply of hot water can be calculated as:
QHW = 365 ⋅ P ⋅ Qperson
The storage size for a four-person household with an average consumption of 45
litres or 1.8 kWh of hot water per day would thus be:
Vstorage = 2 ⋅ 4 ⋅ 45 litres = 360 litres

Typical storage sizes are 300 or 400 litres. A 400-litre storage would be more than
adequate in this case. A 300-litre storage falls somewhat below the calculated
requirement but is still sufficient. The annual heating requirement is:
QHW = 365 ⋅ 4 ⋅1.8 kWh = 2628 kWh.

Collector Size

Once the size of the storage cylinder has been established, the collector size Acollector
is calculated. This calculation also requires the annual figures for radiation Hsolar and
tilt gains forientation (see Section 5.4). With an annual solar fraction sf and an average
system efficiency of 30% with systems using flat-plate collectors, the collector size
Acollector can be calculated as follows:
sf ⋅ QHW
Acollector ≈
0, 3 ⋅ Hsolar ⋅ forientation
In this formula a solar fraction of 0.6 should be chosen for European climates (e.g.
Berlin or London) and 0.9 for very sunny climates such as California.
In the case of the above consumption, the collector size for flat-plate collectors
should be calculated on the basis of a roof in Berlin orientated about 30 degrees
towards the south-southwest and tilted by 30 degrees. The annual solar radiation in
this case amounts to 1000 kWh/(m2 a) and the southern orientation produces tilt

135
6 Solar Thermal Systems – Year-Round Heating from the Sun

gains of forientation = 110% = 1.1. As a result, the required surface for flat-plate col-
lectors is calculated as
0.6 ⋅ 2628 kWh
Acollector ≈ = 4.8 m 2 .
0.3 ⋅1000 kWh
m2
⋅1.1

The results of course depend heavily on the quality of the collectors and can vary
considerably. Some online tools are available to help with system design (see Web
tips). Sophisticated computer programmes are necessary to improve the detailed
planning. Professional firms that specialize in this field should also be able to
provide detailed system designs. In addition to determining the size of the collector
and the storage, their services include the design of other components such as
pumps, controllers and pipes.

■ http://www.valentin.de/onlineberechnung/ Online calculations for solar


solarthermie thermal systems
■ http://desire.htw-berlin.de

6.4.2 Solar Thermal Heating as Support Heating


A large collector surface is needed if a solar thermal system is to provide support
heating in addition to domestic hot water. In contrast to the supply of hot water,
this option requires optimal building insulation to enable the sun to provide a larger
share of the heat requirement. Whereas hot water use is relatively constant through-
out the year, heating needs are concentrated in the winter months. However, the
yield from solar collectors is low in the winter. Therefore, solar thermal systems
that supply support heating are usually designed so that, in addition to hot water,
they can cover only a portion of the heating required during the transitional period
from March to October. In winter conventional heating systems essentially provide
the heating required (Figure 6.20).
The size of the collector surface and the storage also affect the degree of solar
coverage, and thus the proportion of heat covered by the sun. This also reduces the
share a conventional heating system provides. If a fossil system fired with oil or gas
is used, then the carbon dioxide emissions drop in accordance with the size of the
solar system. However, a very large system also produces a higher surplus that
cannot be used. Therefore, as a rule, large systems are less economical than small
ones. So when it comes to design, one has to consider whether the priority is
maximum input from the sun or economic viability.
The following two design variants provide the basis for an outline design in Central
European climates:
Variant 1: Small system for good efficiency
■ Collector surface with flat-plate collectors: 0.8 m2 per 10 m2 living area

136
6.4 Planning and Design

2250
kWh
therm al heat boiler
2000
drinking water boiler
therm al heat solar
1750
drinking water solar

1500

1250

1000

750

500

250

0
Jan. Feb. March April May June July Aug. Sept. Oct. Nov. Dec.

Figure 6.20 Typical progression of thermal heat and hot water requirements in Germany
and proportion of solar system versus conventional heating based on the requirements
of an old building with a total solar coverage of 20%.

■ Collector surface with vacuum-tube collectors: 0.6 m2 per 10 m2 living area


■ Storage size: at least 50 litres per m2 of collector surface
Variant 2: Medium-sized system for higher share of solar coverage
■ Collector surface with flat-plate collectors: 1.6 m2 per 10 m2 living area
■ Collector surface with vacuum-tube collectors: 1.2 m2 per 10 m2 living area
■ Storage size: 100 litres per m2 of collector surface
An optimal design would of course also take into account the actual heat require-
ments. The difference in requirements between an old building and an energy-saving
three-litre house is considerable. Table 6.1 shows simulation results for optimal
systems that were dimensioned according to the outline design described.

Table 6.1 Solar coverage.


Old Standard Three-litre Passive
building new-build house house
Heat requirement for hot water in kWh 2700 2700 2700 2700
Heating requirement in kWh 25 000 11 500 3900 1950
Solar coverage Variant 1 (small system) 13% 22% 40% 51%
Solar coverage Variant 2 22% 36% 57% 68%
(medium-sized system)
Assumes following: Berlin location, orientation 30 ° south without shade, 130 m2 living area, optimal flat-plate collector with combination storage.

Although the collector with the medium-sized system is double the size of the small
system and the storage is four times larger, this does not mean that the solar cover-

137
6 Solar Thermal Systems – Year-Round Heating from the Sun

age is twice as high. The insulation standard of a building has considerably more
influence on the rate of solar coverage than these two elements. Optimal insulation
options should be considered when there is an interest in increasing the amount of
solar coverage as a way of contributing towards climate protection.

From Thinking about a Solar Thermal System on the Roof to


Installing a System

■ Determine orientation and angle of roof.


Is the roof suitable in terms of orientation and tilt angle?
Recommendation: a minimum of 95% according to Figure 5.15.
Does the roof have too much shade?
■ Decide whether only water for domestic use should be heated or whether solar
heating or cooling support is also desired.
■ Implement outline design following rule of thumb.
■ Possibly implement detailed design by hand or using simulation tools.
■ Apply for appropriate approvals for listed buildings.
■ Request quotations.
■ Have professional company provide additional details for design.
■ Apply for subsidies or low-interest loans.
■ Award contract for the work.

6.5 Economics

Depending on type and design, between 200 and 350 euros per square metre should
be estimated for flat-plate collectors and between 400 and 600 euros per square
metre for vacuum-tube collectors in Central Europe. A 300-litre heat storage cylin-
der costs around 700 to 1100 euros. Installing a solar thermal system can be espe-
cially cost-effective in new buildings or when an existing hot water storage cylinder
is old and has to be replaced anyway. The costs for a solar thermal system fall within
a wide range. A system purely for domestic hot water with four square-metre flat-
plate collectors and a 300-litre hot water storage cylinder can be acquired for as
little as 2000 euros, not including installation. The average cost of a European
system for a four-person household excluding installation is between 3000 and 3400
euros; a system including installation is around 5000 euros. Depending on the size
of the collector, the cost of a system that provides support heating can be double or
even higher. Government grants are sometimes available to help cover the costs.
Even if the investment costs for a solar thermal system are known, getting a handle
on the economics is difficult compared to photovoltaics. The output of a photo-
voltaic system can be tracked accurately through an electricity meter and the state-

138
6.6 Ecology

ment for the fed-in electricity will show in euros and cents whether a system is
living up to the planners’ promises.
The output of a solar thermal system can also be monitored with a heat volume meter.
However, in practice these are hardly ever used because of cost. In very sunny coun-
tries a solar thermal system can cover total requirements. The cost-effectiveness is
also usually high in these cases. In colder climates solar thermal systems are almost
always supplemented by conventional heating systems, which make up for what the
lack of sunshine cannot cover. No direct compensation is given for the heat output
yield of a solar thermal system. A system only pays for itself indirectly through the
savings in the fuel costs of the heating system. It is mainly the movement of fuel prices
that establishes whether a system is paying for itself and at which rate. The higher
fuel prices climb, the faster a solar thermal system will pay for itself. If, on the other
hand, fuel prices fall, the economic viability of a solar system is less favourable.
Figure 6.21 shows the payback periods for a typical solar thermal system for heating
domestic water. This is the time it takes for the fuel costs a system has saved to
break even with the investment costs. If hot water is heated electrically for around
0.2 euro/kWh, the system will be amortized very quickly. The amortization becomes
more difficult with particularly high-quality and expensive systems when fuel prices
are low. In this case public grant programmes could be available to improve the
economy of these systems.

20
payback period in years

15

10

investment costs in €
2000 2500 3000
5
3500 4000 4500
5000 5500 6000

0
0.05 0.1 0.15 0.2 0.25
fuel price in €/kWh

Figure 6.21 Payback periods for a solar thermal domestic hot water system with backup
heating system (without interest rate effects and price increases; fuel savings:
2000 kWh/a; annual operating costs: 2% of investment costs).

6.6 Ecology

Solar thermal systems are among the most environmentally friendly renewable
energy systems available. When they are used, they normally save on fossil
fuels, such as oil and natural gas, and thereby contribute actively towards climate

139
6 Solar Thermal Systems – Year-Round Heating from the Sun

protection. The collectors are mostly integrated into buildings and, consequently,
do not require any extra land. The materials used in solar thermal systems –
such as glass, copper and plastic – are for the most part just ordinary materials
commonly used in standard construction. Materials that are a problem for the
environment, such as PU-foam and PVC, are used in some solar collector systems,
but most manufacturers of collectors consciously avoid them as a matter of
principle.
Energy is needed to manufacture solar thermal systems. In Central Europe, it takes
between six months and two years before a solar thermal system delivers the same
amount of energy that was used in its production. This period is shorter in countries
with a lot of sunshine. Many manufacturers place a great deal of emphasis on
environmental protection and renewable energies during the production of thermal
solar systems. The zero-emissions Solvis factory in Braunschweig, Germany, is an
example.
Some attention needs to be paid to the size of electrical pumps in solar systems.
These pumps require electric energy to operate. However, this requirement for
auxiliary energy is usually smaller by many orders of magnitude than the energy
saved by a solar system. Photovoltaic systems can be used to enable the auxiliary
electric energy required by solar thermal systems to be covered directly by the sun.
Many solar thermal systems also use chemicals such as antifreeze or cooling agents.
Typical antifreeze products like Tyfocor L have a minimal effect on water quality,
and, therefore, are largely safe for the environment.
Special protective measures are required if ammonia is used for solar cooling in
absorption-refrigerating machines. Ammonia is toxic and dangerous to the environ-
ment. Ammonia that escapes can bind to water. Lithium-bromide is also harmful to
human health, but less so than ammonia.

6.7 Solar Thermal Markets

In terms of world markets for solar thermal collectors, one country puts all the rest
to shame: China. Around 93 million square metres of collectors were installed in
the country in 2006. These collectors deliver about 65.1 gigawatts of heat power.
The Chinese collector market comprises more than 60% of the entire world market
(Figure 6.22).
One of the main reasons for the booming solar market in China is the poor supply
of electricity and gas in rural regions. Seventy-five percent of the 1.3 billion Chinese
live in the country. In these rural areas solar thermal energy is often one of the least
expensive alternatives for providing hot water, and sometimes the only one. With
its widespread use of solar thermal energy, China has developed into a market leader
in collector manufacturing. More than 1000 companies now produce and distribute
solar collectors in China. Around 150 000 people were employed in the solar thermal
energy (European Solar Industry Federation ESTIF, 2003) sector in the year 2000.
In contrast to many other countries, China specializes in highly efficient vacuum-

140
6.7 Solar Thermal Markets

China
93 km²

others
15.4 km²

USA
2.3 km²
Austria
2.7 km²
Greece
3.3 km²
Israel
4.9 km²
Japan
6.8 km²
Germany
8.1 km² Turkey
9.5 km²

Figure 6.22 Total installed collector area in different countries without swimming pool
absorbers. Status 2006. Data: (IEA Solar Heating and Cooling Programme SHC, 2008).

tube collectors rather than in simple flat-plate collectors. Due to the large number
of units produced for the Chinese market, Chinese suppliers are very competitive
internationally. Most of the vacuum-tube collectors sold in the world come from
China.
In the European Union, Germany dominates the solar collector market. Countries
with a lot of sunshine, such as Spain, Portugal and France, are still developing
countries as far as solar thermal systems are concerned.
If one considers solar thermal collector area installed per head, small countries have
the edge. In Cyprus almost every house has a solar thermal system on it. In 2008
this Mediterranean island had around 693 000 m2 of collector area distributed among
778 000 inhabitants. There are nearly 900 m2 of solar collectors per 1000 Cypriots.
In Austria, in comparison, this figure is 388 m2, and for Germany it is less than 135 m2
per 1000 inhabitants. Sunny Spain fares even more poorly with just 30 m2 per 1000
inhabitants, and in the United Kingdom the figure is a paltry 6 m2. Local market
conditions and acceptance by the population both have considerably more influence
on the amount of installed collector area than the supply of solar radiation.
Germany is an example of how political conditions can affect markets. Until 2001
there was continuous market growth. In 2002 the market suffered a setback because
of changes in incentive conditions and associated market insecurity (Figure 6.23),
but has since recovered.
Altogether more than one million solar thermal systems with a total area of 11
million square metres were installed at the end of 2008. This constitutes more than
40% of the 27.2 million square metres in the European Union. In the same year the
solar thermal sector in Germany accounted for 25 000 jobs. The solar systems are

141
6 Solar Thermal Systems – Year-Round Heating from the Sun

2 200 000

2 000 000
1 800 000
1 600 000
1 400 000
1 200 000
1 000 000
800 000
600 000
400 000
200 000
0
1990

1991

1992

1993

1994

1995

1996

1997

1998

1999

2000

2001

2002

2003

2004

2005

2006

2007

2008
Figure 6.23 Annual newly installed collector area in Germany. Data: (Bundesverband
Solarwirtschaft, 2009).

saving the emission of around one million tons of carbon dioxide each year. Although
this is an impressive number, it still falls far short of saving the climate.

6.8 Outlook and Development Potential

During the past fifteen years the solar thermal market in Europe has increased more
than tenfold. The European Solar Thermal Industry Federation (ESTIF) has formu-
lated an ambitious target for Europe. According to ESTIF about one square metre
of installed collector area per capita could be possible until 2020 and more than
three square metres in the long term (Figure 6.24). This would be 60 times more
than in 2008. The number of people employed in the European solar thermal sector
could then increase to hundreds of thousands. The potential is considerably higher

Figure 6.24 Many roofs still have space for solar collectors. Photos: www.wagner-solar.com.

142
6.8 Outlook and Development Potential

in the very sunny countries of the world. Here the solar thermal share could reach
percentages in double figures. In addition to those aimed at heat production, systems
that provide solar cooling will gain large market shares and in the long term displace
conventional cooling systems. Due to the growth in the number of installations, it
is anticipated that the costs of solar thermal systems will continue to drop slightly.
However, it is not expected that the reductions in price will be quite as high as with
photovoltaic systems.

143
7
7 Solar Power Plants – Even More
Energy from the Sun
When we think of power plants, we usually imagine large central complexes with
cooling towers and enormous chimneys emitting clouds of smoke. In terms of the
concept itself, a power plant is merely a technical installation that converts a par-
ticular energy supply into electricity. A solar power plant produces electric energy
from solar radiation. We have already learned about one type of solar power plant
– photovoltaic power plants. In principle, a small photovoltaic system sited on a
family home already constitutes a solar power plant. Because many people envisage
a power plant as something much larger than a few square metres of solar modules,
this chapter is devoted to really large-scale systems that generate electricity from
sunlight.
Even photovoltaic systems can reach a size that easily competes with conventional
power plants. Year after year new photovoltaic power plants have been exceeding
records for size. Photovoltaic systems with 5 megawatts of power and module areas
of around 40 000 m2 have already been installed on the roofs of industrial buildings.
However, existing roofs are usually not appropriate for higher output levels. If 50,
100 or even 1000 megawatts are to be installed in one location, this can only be
done with open space systems mounted directly in the ground. Large photovoltaic
power plants are technically very similar to the small systems on single-family
houses, except that they are larger in scale. Thus this chapter only deals with large-
scale solar power plants.
Concentrators can be used to increase the intensity of solar radiation. Photovoltaic
cells with very high efficiency then convert the concentrated sunlight into elec-
tricity, as this chapter will explain. In addition to photovoltaics, other technologies
for generating power from sunlight also exist. Solar thermal power plants convert
solar radiation first into heat and then into electric energy. A number of different
interesting and very promising technologies are available and are particularly
suitable for the very sunny regions on earth because they need direct sunlight,

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

144
7.1 Concentration on the Sun

which is far less abundant in temperate climates such as that of Northern


Europe.

7.1 Concentration on the Sun

Most of us remember as children trying to set fire to a sheet of paper or a piece of


wood using a magnifying glass. Even this simple experiment makes us aware of the
energy that concentrated solar radiation can produce. Theoretically, sunlight can be
concentrated onto earth by the factor 46 211 and, consequently, at its deepest point
reach temperatures of 5500 °C. In practice, concentration factors of over 10 000 and
temperatures of well over 1000 °C have been reached. Large holes can easily be
melted into steel plates using concentrated solar radiation in solar ovens. Solar ovens
are also suitable for testing materials at high temperatures. Figure 7.1 shows a solar
oven near Almería in Spain, which reaches an output of 60 kilowatts with full sun
radiation.

Figure 7.1 Solar furnace near Almería in Spain. Large tracking mirrors direct the
sunlight towards a convex mirror (top right) in the interior of a building.

For cost reasons, glass lens systems are normally ruled out for use as concentrators
in large-scale technical applications. The reflector, which concentrates sunlight onto
a focal point or at the focus, normally has the form of a parabola. Due to their long
useful life, glass mirrors have proven to be reliable in practical application. The
reflector has to be tracked so that the sunlight always comes in at a vertical angle.

145
7 Solar Power Plants – Even More Energy from the Sun

A distinction is mainly made between single-axis and dual-axis tracking systems.


Single-axis tracking systems concentrate the sunlight onto an absorber pipe at the
focus, whereas dual-axis tracking systems concentrate it onto a central absorber
close to the focus. The tracking can occur either over a sensor, which catches the
optimal orientation to the sun, or through a computer that calculates the sun’s
position.

Archimedes – Inventor of the Convex Mirror?


The Greek mathematician Archimedes of Syracuse, who was born in 285
BC, is considered the inventor of the convex mirror. Using convex mirrors,
Archimedes is said to have set fire to enemy ships while his country was at war with
Rome. The truth of this legend has been hotly debated for the last 300 years, but it is
now considered to have been unlikely. It would have taken very large and very precisely
made mirrors, focusing on the ships over a very long period, to light the fires. Even if
the legend is true, it did not do Archimedes much good. He was killed by the Romans
when they captured Syracuse in 212 BC.

Parabolic trough collectors, which concentrate sunlight onto an absorber pipe, are
normally used as line concentrators (Figure 7.2 left). If the concentrator is distrib-
uted over multiple mirrors each individually lined up in an optimal position towards
the absorber pipe, this is called a Fresnel collector (Figure 7.2 right).

Figure 7.2 Single-axis tracked reflectors for line concentrators.

Convex mirrors are used for concentration onto a focal point (Figure 7.3 left).
Distributed mirrors that are individually adjusted towards the sun can also concen-
trate the sun’s radiation onto a central absorber (Figure 7.3 right).

146
7.2 Solar Power Plants

Figure 7.3 Dual-axis tracked reflectors for point concentrators.

7.2 Solar Power Plants

7.2.1 Parabolic Trough Power Plants


As the name indicates, in parabolic trough power plants large trough-shaped para-
bolic mirrors concentrate sunlight onto a focal point. The collectors are erected next
to each other in a row several hundred metres long (Figure 7.4). Many parallel rows
in turn form an entire solar collector field.

Figure 7.4 View of the Kramer Junction parabolic trough power plant in California (USA).
Photo: Gregory Kolb, SANDIA.

147
7 Solar Power Plants – Even More Energy from the Sun

The individual collectors rotate on their longitudinal axis and in this way follow the
course of the sun. The mirrors concentrate the sunlight more than 80-fold at the
focal point onto an absorber pipe. This is embedded in an evacuated glass casing
in order to reduce heat loss. A special selective coating on the absorber pipe reduces
the heat radiated from the pipe surface. With conventional systems, a special thermal
oil flows through the pipe, which heats up to temperatures of about 400 °C as a
result of the solar radiation. The heat is transmitted over heat exchangers to a water
steam cycle, vaporized and overheated again. The steam drives a turbine and a
generator, which produces electric power. Behind the turbine it condenses again
into water and through a pump again enters into the cycle (Figure 7.5). The principle
of producing electricity using steam turbines is called the Clausius-Rankine process,
named after its inventors. This process is also used in classic steam power plants,
such as coal-fired plants.

Figure 7.5 Parabolic trough power plant with thermal storage.

During periods of bad weather or at night a parallel burner can also be used to
operate the water steam cycle. In contrast to photovoltaics, this guarantees a daily
output of power. It also increases the attractiveness of and planning security in the
public electricity supply. For totally carbon-free plant operation, either biomass or
renewably produced hydrogen can be used as a supplementary fuel, or the burner
can be eliminated entirely. Instead a thermal storage tank can be integrated. The
solar field heats up the storage during the day using excessive heat. At night and
during periods of bad weather the storage feeds the water steam cycle (Figure 7.6).
The storage must be designed to handle temperatures above 300 °C. Molten salt is
suitable as a storage medium for this temperature range.
The development of solar thermal parabolic trough power plants dates back to
1906. In the USA and near the Egyptian city of Cairo – at the time still under British
rule – research facilities were set up and the first tests were successful. Based on

148
7.2 Solar Power Plants

250

200
heat to
power in MW the storage
150
secured power

100

heat from electricity heat from


50 the storage direct from sun the storage

0
00:00
01:00
02:00
03:00
04:00
05:00
06:00
07:00
08:00
09:00
10:00
11:00
12:00
13:00
14:00
15:00
16:00
17:00
18:00
19:00
20:00
21:00
22:00
23:00
00:00
Figure 7.6 Solar thermal power plants with thermal storage can provide guaranteed
output around the clock.

appearance, these trough facilities were incredibly similar to those used today.
However, problems with materials and other technical difficulties ended this first
attempt at large-scale technical generation of solar electric power in 1914, shortly
before the start of the First World War (Mener, 1998).
In 1978 the foundations were laid in the USA for a resurrection of this technology.
The Public Utility Regulatory Policies Act obligated American public electricity
companies to buy energy from independent producers at clearly defined costs. After
the surge in energy costs in the wake of the oil crisis, the electricity provider
Southern California Edison (SCE) offered long-term feed-in conditions. The intro-
duction of favourable tax advantages finally made the construction of plants worth-
while financially. The LUZ Company, which was founded in 1979, negotiated a
30-year contract for the feed-in of solar energy with SCE in 1983. In 1984 the first
solar thermal parabolic trough power plant was built in the Mojave Desert in
California. By 1991 a total of nine SEGS (Solar Electric Generation Systems) power
plants with 354 megawatts of electric output were installed in an area of more than
7 km2. These power plants feed around 800 million kilowatt hours into the grid each
year, enough to satisfy the requirements of 60 000 Americans. Eight power plants
can also be operated with fossil fuels, thus enabling them to supply electricity at
night or during periods of bad weather. With these plants the annual natural gas
portion of the thermal energy supplied is legally set at 25%. The total investment
for the plants was more than US$1.2 billion.
In the mid-1980s energy prices fell dramatically again. Then the tax exemptions
also ran out at the end of 1990, and the LUZ Company went bankrupt even before
construction could start on the tenth power plant.
A long lean patch then followed for the planners of solar thermal power plants. It
lasted until 2006, when building work was started on new parabolic trough plants
in Nevada in the USA and near Guadix in Spain.

149
7 Solar Power Plants – Even More Energy from the Sun

Technical advances made in the meantime have helped to increase efficiency and
reduce costs. One option was the use of direct solar steam. With this process water
is vaporized at a high pressure by the collectors and heated up to 500 °C. This steam
is led directly into a turbine, thereby rendering thermo oil and heat exchangers
superfluous.

Temperature and Efficiency


The Achilles heel of solar thermal power plants is not the solar collector that
concentrates the sunlight. Collectors easily achieve efficiency of over 70%.
However, most of the valuable solar heat is lost during its conversion into electricity.
The steam turbines used in this process in solar power plants barely manage to achieve
35% efficiency. In other words, 65% of the heat gained from the sun returns unused as
waste heat into the environment. The efficiency of steam turbines results directly from
the temperature difference of the steam between entry into and exit from the turbines.
The exit temperature depends on the cooling and even at best is only minimally below
the ambient temperature. With parabolic trough power plants, the entry temperature,
contingent on the thermo oil, is currently just below 400 °C. With a temperature increase
to 500 degrees or higher, the efficiency of turbines could easily reach 40%. But this is
the maximum even for steam turbines. Combined gas and steam turbines, which operate
at temperatures of over 1000 °C, achieve efficiency of up to 60%. Gas and steam turbines
with high efficiency can be used, for example, in solar tower power plants.
One question that is often asked is whether simple pipe collectors can be used to generate
electricity from general purpose water. In principle, this would be possible. However,
due to the extremely low temperatures, efficiency would be too minimal to make this
approach economically viable. Even the waste heat in the environment would have
limited use. Solar power plants are normally situated in hot, sunny regions. However, a
demand for gigantic quantities of low-temperature heat does not really exist in these
areas.

7.2.2 Solar Tower Power Plants


With solar tower power plants, several hundred or even thousand rotating mirrors
are arranged around a tower. Called heliostats, these mirrors are individually con-
trolled by computer to track the movement of the sun and are orientated towards
the top of the tower. They must be orientated precisely within a fraction of a degree
so that the reflected sunlight actually reaches the focal point. A receiver is located
here with an absorber, which, due to the highly concentrated sunlight, heats up to
temperatures of over 1000 degrees. Air or molten salt transports the heat. A gas or
steam turbine that drives a generator ultimately converts the heat into electric
energy.
With the tower concept with open volumetric receivers (Figure 7.7), air from the
environment is sucked by a blower through a receiver towards which the heliostats
are orientated. The materials used for the receiver are wire mesh, ceramic foam or
metallic or ceramic honeycomb structures. The receiver is heated up by the solar
radiation and transfers the heat to the sucked-through ambient air. The sucked-in
air cools the front side of the receiver. Very high temperatures only develop on the

150
7.2 Solar Power Plants

Figure 7.7 Solar tower power plant with open air receiver.

inside of the receiver. This so-called volumetric effect reduces radiant heat losses.
The air, which is heated to temperatures between 650 °C and 850 °C, enters a waste-
heat boiler that vaporizes and overheats the water, thereby driving a steam turbine
cycle. If required, this variant of power plant can be fired over a channel burner
using other fuels.
A more advanced version of the tower concept that uses a pressure receiver is now
offering promising possibilities for the future (Figure 7.8). With this concept, the

Figure 7.8 Solar tower power plant with pressurized air receiver.

151
7 Solar Power Plants – Even More Energy from the Sun

concentrated sunlight heats the air in a volumetric pressure receiver at about 15 bars
to temperatures up to 1100 °C. A transparent quartz-glass dome separates the
absorber from the environment. The hot air then drives a gas turbine. The waste
heat of the turbine then finally drives the downstream steam turbine process. The
first prototype has shown that this technology functions successfully.
With the combined gas and steam turbine process the efficiency of the conversion
from heat to electric energy can be increased from about 35% to more than 50%
with a pure steam turbine process. As a result, total efficiencies of more than 20%
are possible with this type of conversion of solar radiation into electricity. These
prospects justify the additional complexity and cost of receiver technology.
In contrast to parabolic trough power plants, there has not been much experience
with commercial plants in the solar tower plant area. Research facilities that are
optimizing system components and testing new components currently exist in
Almería (Spain) (see Figure 7.9), Daggett (USA) and Rehovot (Israel).

Figure 7.9 Research site for a solar tower power plant at Plataforma Solar de Almería
(Spain).

The first commercial solar tower power plant to be put into operation was the
11-megawatt PS10 tower plant near Seville in Spain in 2006. However, instead of
heating up air, the receiver of this plant vaporizes water. Due to the low temperatures
the efficiency of this power plant is still relatively low. In 2006 construction started
on the 20-megawatt PS20 tower plant, also near Seville, and other solar tower plants
are planned there.
Before it can be launched successfully in the market, the open air receiver technol-
ogy developed in Germany first has to prove its suitability for practical use. This is
currently being tested at a newly built solar tower plant in Jülich, Germany. With
1.5 megawatts of power this demonstration plant is considerably smaller than the
commercial Spanish plants. The target in this effort is not the German power plant
market but to promote the export of German technology to the sunny countries of
the world.

152
7.2 Solar Power Plants

7.2.3 Dish-Stirling Power Plants


Whereas trough and tower power plants are only economically viable in large-scale
applications of many megawatts, the so-called Dish-Stirling systems can also be
used in smaller units – for example, to supply remote villages or towns. With Dish-
Stirling systems a convex mirror in the form of a large dish concentrates the light
onto a focal point. To ensure that the light is concentrated as strongly as possible,
the mirror is dual-axis tracked very precisely towards the sun.
A receiver is sited at the focal point. This receiver transfers the heat to the actual
heart of the system: the Stirling engine. This engine converts the heat into kinetic
energy and drives a generator that ultimately produces electric energy.
A Stirling engine can be driven not only by the heat of the sun but also through
combustion heat. In combination with a biogas burner these plants can also generate
electricity at night or during periods of bad weather. And the use of biogas also
makes them carbon-neutral.
Some prototypes of pure solar systems have been built in Saudi Arabia, the USA
and Spain (Figure 7.10). Compared to tower and trough power plants, the price per
kilowatt hour with Dish-Stirling systems is still relatively high. The costs would be
reduced considerably if large unit quantities could be mass produced.

Figure 7.10 Prototype of a 10-kW Dish-Stirling system near Almería in Spain.

7.2.4 Solar Chimney Power Plants


There is a big difference between solar chimney power plants and the thermal plants
described earlier. While solar thermal power plants work by concentrating sunlight,
solar chimney power plants function through the heating of air. The collector field
is formed by a large flat area that is covered by a glass or plastic roof. A high
chimney is located in the middle of the area, and the collector roof rises gently in

153
7 Solar Power Plants – Even More Energy from the Sun

the direction of the chimney. The air is able to flow in unencumbered at the sides
of the enormous roof. The sun warms the air under the glass roof. This air then rises
upward, follows the gentle slope of the roof and then flows at great speed through
the chimney. The airflow in the chimney then drives wind turbines that generate
electric power over a generator.
The ground under the glass roof can store heat so that the power plant is still able
to deliver power even after the sun has set. If hoses filled with water are laid in the
ground, enough heat can be stored to enable the plant to provide electric power
around the clock.
At the beginning of the 1980s a small demonstration plant with a rated power output
of 50 kilowatts was built near Manzanares in Spain. The collector roof of this plant
had an average diameter of 122 m and an average height of 1.85 m. The chimney
was 195 m high and had a diameter of 5 m. This plant was dismantled in 1988 after
a storm knocked the chimney down. However, all the planned tests had been com-
pleted and the research plant lived up to expectations. It was the first successful
demonstration of a solar chimney power plant.
Because the efficiency of solar chimney plants compared to other techniques is very
low, these plants require large areas of land. Furthermore, the efficiency increases
in line with the height of the tower. Therefore, to be economically viable plants
must be of a certain minimum size. New power plant projects are currently being
discussed in Australia, for example. Consideration is being given to a large-scale
200-megawatt plant with a tower height of 1000 m, a tower diameter of 180 m and
a collector diameter of 6000 m. It is difficult to predict whether the building of new
large-scale power plants will be financially viable. But based on a long-term view,
solar chimney power plants in the desert regions of the world have the potential to
be financially competitive compared to conventional plants (Figure 7.11).

Figure 7.11 Computer animation of a solar chimney power plant park. The towers can
also be used as viewing platforms. Illustration: Schlaich Bergermann Solar, Stuttgart.

154
7.2 Solar Power Plants

7.2.5 Concentrating Photovoltaic Power Plants


Photovoltaic cells can also be operated by concentrating sunlight. The main point
of this is that the concentration saves on valuable solar cell material. If sunlight is
concentrated by a factor of 500, the size of the solar cell can then also be reduced
by a factor of 500. The cost of the solar cell therefore becomes much less of an
issue. This means materials that ordinarily would be too expensive without solar
concentration could also be used. Concentrator cells therefore usually have higher
efficiency than conventional photovoltaic modules.
There are many options for the concentration: for example, concentrator cells can
be mounted in the focal point of parabolic troughs or convex mirrors. One of the
main problems with this is efficient cooling, because, in addition to the electric
energy of the solar cells, a large quantity of waste heat is produced. Flatcon technol-
ogy takes a different approach. A flat Fresnel lens concentrates the sunlight onto a
concentrator cell that is only a few square millimetres in size (Figure 7.12 below
right). A copper plate on the back of the cell radiates the heat that accumulates
extensively towards the back. A concentrator module comprises a large number of
parallel cells. Many modules are then mounted together on a tracking device that
orientates the modules optimally towards the sunlight.

Figure 7.12 Photovoltaic system with concentrator cells. Photo/graphics: Concentrix Solar
GmbH.

7.2.6 Solar Chemistry


In addition to providing process heat or generating electricity, concentrating solar
thermal energy can also be used for testing materials or in solar-chemical systems.
For instance, there is a large solar furnace in the French town of Odeillo where a
large number of small mirrors are mounted on a hillside. These mirrors reflect the
sunlight to a convex mirror 54 m in diameter, creating temperatures of 4000 °C in
the solar furnace, which is used for research and industrial processes. Solar furnaces
have also been built in Almería, Spain, and Cologne, Germany.

155
7 Solar Power Plants – Even More Energy from the Sun

In addition to producing chemicals at high temperatures, solar thermal energy can


also be used to produce hydrogen. This does not require a circuitous route that starts
with generating electricity, followed by electrolysis. At high temperatures hydrogen
can be produced through a solar-chemical process. For example, the chemical
system could be in the receiver of a solar tower. Hydrogen is treated as an important
energy source, particularly in the transportation sector and in fuel cells. If a hydro-
gen economy ever becomes a reality, concentrating solar-chemical systems could
play a major role in hydrogen production.

7.3 Planning and Design

As a rule, solar-thermal power plants are classic plants. Due to their size, they cost
millions of euros. Solar plants are almost always planned and built by large corpora-
tions or industrial companies. The design of these plants is usually very complex.
It takes teams of engineers a long time to do the detailed planning. One of the main
goals is to optimize power plants from an economic point of view.
Private individuals are not likely to have any involvement in the planning of solar
power plants, unlike the situation with small photovoltaic systems and solar thermal
systems that heat tap water or provide supplemental heating. However, the rise in
the number of solar power plants is presenting many opportunities for investment
in this area. Therefore, it is worth taking a quick look at the planning aspects.
Because concentrating solar power plants suffer a major reduction in efficiency in
the partial load area, the aim should be to build them mainly in countries that get a
lot of sunshine. The regions currently of interest are those with an annual total global
radiation of at least 1800 kilowatt hours per square metre (kWh/m2). The optimal
values are clearly those over 2000 kWh/m2. These areas appear in red or pink in
Figure 7.13.

640.. 900
900.. 1050
1050.. 1200
1200.. 1350
1350.. 1500
1500.. 1700
1700.. 1900
1900.. 2100
2100.. 2300
> 2300

Figure 7.13 World map with annual totals for solar global radiation in Wh/m2. Source: Meteotest, www.
meteonorm.com.

156
7.3 Planning and Design

7.3.1 Concentrating Solar Thermal Power Plants


Concentrating solar power plants can only utilize the direct irradiance portion of the
sun. This portion of irradiance can be reflected by mirrors and ultimately concen-
trated. For non-concentrating solar systems the global irradiation intensity, thus the
sum of direct and non-directional diffuse sunlight, is important. The efficiency of
these systems is therefore related to the global irradiation intensity. With concentrat-
ing systems the direct-normal radiation intensity on a surface orientated vertically
towards the sun serves as a reference quantity (see Figure 7.14). In the technical
jargon this is abbreviated to DNI, which stands for ‘direct normal irradiance’. The
DNI values can lie somewhat below but also somewhat above the values for global
radiation.

Figure 7.14 Differentiation of types of solar radiation.

Annual Electricity Production of Solar Power Plants

The annual output yield of a concentrating solar thermal system can be estimated
using the average efficiency η, the annual total of solar direct-normal irradiance
DNI and the aperture area A of the mirrors:
Eelectrical = η ⋅ A ⋅ DNI .
With an efficiency η of 15% = 0.15, a solar thermal parabolic trough system with
an overall aperture area of 500 000 m2 at a Spanish location and a DNI of 2200 kWh/
(m2 a) achieves an annual yield of:
kWh kWh
E = 0.15 ⋅ 500 000 m 2 ⋅ 2200 = 165 million .
m2 a a
This is sufficient to cover the electricity needs of around 50 000 Spanish
households.

157
7 Solar Power Plants – Even More Energy from the Sun

Another interesting aspect of planning is how the land is used. So that the mirrors
do not throw shadows on one another, they can only be installed on about one-third
of the available land area. The land area for a concentrating solar power plant must
therefore be at least three times bigger than the surface area of the mirrors.
The mirror area should be coordinated optimally with the rest of the plant. If it is
too small, the plant will constantly be running in partial-load operation. As a result,
both efficiency and yield will drop. If the size of the mirror area is too large, a higher
amount of solar radiation will fall onto the collectors than can ultimately be con-
verted into electricity. In this case some of the mirrors have to be turned away from
the sun.
The efficiency depends on the technology used. Dish-Stirling power plants and solar
tower plants with pressure receivers can reach efficiencies of 20% or more. Solar
tower or trough plants with steam turbines are currently at about 15% efficiency.
An increase in DNI irradiance values also increases efficiency as a power plant will
then be running for shorter periods in partial-load operation. In Germany efficiency
in the order of 10% is all that can be expected from its solar thermal power plants
– and this is based on moderate DNI irradiance values of about 1000 kWh/m2 and
by year. The annual yield from concentrating solar power plants in places like Spain
would be roughly three times higher than in Germany or Great Britain.

7.3.2 Solar Chimney Power Plants


With solar chimney power plants the annual yield is calculated on an analogous
basis. Instead of the DNI, the global radiation is used because these plants can also
use diffuse sunlight. The total efficiency is only about 1% – and then only if the
tower is more than 1000 m high. Thus the efficiency of a solar chimney power plant
is directly correlated to the height of the tower. At half the optimal tower height the
efficiency is also halved. The collector area would have to be doubled in size to
achieve the same yield. In contrast to a concentrating power plant, a solar chimney
plant can use the entire land area. No unused gaps are necessary to prevent shading.

7.3.3 Concentrating Photovoltaic Power Plants


The annual yield from concentrating photovoltaic power plants is also calculated
analogously. Compared to non-concentrating photovoltaic plants, an efficiency
of more than 20% is possible. As concentrating photovoltaic plants can only
use the direct-normal irradiance portion, the yield in sunny countries rises
overproportionally.

7.4 Economics

Solar thermal plants can currently generate electric power more cost-effectively than
photovoltaic plants – at least in regions with high solar irradiation. In Germany at
best the same electricity generation costs can be expected from both solar thermal

158
7.4 Economics

and photovoltaic plants. As a result, the operation of solar thermal power plants is
less feasible for economic reasons (Figure 7.15).

Figure 7.15 In sunny regions solar thermal power plants (right) can currently produce electricity more
cheaply than photovoltaic power plants (left).

In Spain, the cost of generating electricity from parabolic trough and solar tower
power plants in 2008 amounted to about 20 cents per kilowatt hour; at top locations
in North Africa the cost was about 15 cents per kilowatt hour. So, although these
power plants are less expensive than those using photovoltaics, they are consider-
ably more expensive than conventional fossil or nuclear plants. As with photovolta-
ics, costs are expected to fall considerably when these plants start to be built on a
large scale. Costs are already dropping fast as a result of the many new solar trough
and tower plants being built. Within a decade solar thermal power plants could
already be more economical than conventional plants.
In the case of solar chimney and Dish-Stirling power plants, there are no signs yet
of a comparable rapid market development. Different studies show that these plants
also have the potential in the long term to be competitive in electricity generation.
It is also calculated that concentrating photovoltaics plants will be able to generate
electricity at a lower cost than normal photovoltaic plants in locations with high
solar irradiation. Various suppliers have developed interesting systems over the past
few years, but concentrating systems still have to prove their viability.
The advantage of photovoltaic systems compared to solar thermal systems is modu-
larity. Photovoltaic systems can be built in any size wanted – ranging from milli-
watts all the way to multi-megawatt plants several square kilometres in area. Solar
thermal power plants are always dictated by the minimum output that is needed to
make the plant economically viable. With Dish-Stirling power plants, this is around

159
7 Solar Power Plants – Even More Energy from the Sun

10 kilowatts. All other solar thermal power plants should have a minimum output
of 10 megawatts. Efficiency is increased even further if output levels are between
50 and 200 megawatts. Optimally, a system should be one square kilometre or more
in total size to make it economically viable.
The main advantage of solar thermal systems is the simple integration of thermal
storage. Small storage units used for a few hours only minimally increase the cost
of generating electricity at solar thermal plants. However, they ensure that the power
output of the plants can be guaranteed and, as a result, increase the availability and
the value of electric energy.

7.5 Ecology

Solar power plants without fossil fuel backup facilities do not release any direct
carbon dioxide emissions during operation. When a fossil fuel-fired parallel burner
exists, as is the case with some solar thermal parabolic trough power plants, the
natural gas portion should be kept to an absolute minimum. Due to the low tem-
peratures, the efficiency of solar thermal plants for electricity generation is lower
than that of optimized pure natural gas plants. This aspect is insignificant when
power plants are run purely on solar power. However, if fossil fuels are burnt as
well, the carbon dioxide emissions rise. It is accepted that the addition of fossil fuel
will increase supply reliability and protect against frost. But the fossil portion should
not exceed 10% to ensure effective environmental protection.
Some World Bank projects in developing countries are striving towards an integra-
tion of relatively small parabolic trough collector fields into conventional gas and
steam power plants run on natural gas. Due to technical restrictions, the solar portion
is clearly below 10% there. This type of ISCCS (Integrated Solar Combined-Cycle
System) is not suitable for effective climate protection.
Compared to conventional photovoltaic systems, the production energy required for
thermal solar power plants is lower. Within a year a solar power plant will deliver
more energy than was originally used to produce the plant.
Unlike small photovoltaic and solar thermal systems, solar systems cannot be inte-
grated into buildings. Instead they need large open areas of space comprising many
hectares. Ideally, solar plants should be erected in thinly populated areas where the
land is not needed for other purposes. Fortunately, many sunny regions on earth
have these very characteristics. Very little grows in hot deserts where the sun shines
almost all year round, and the human populations tend to be low. Solar thermal
power plants built on suitable desert sites would easily be able to meet the electricity
demands of the entire planet a hundred times over.
The main problem with thermal systems is the need for cooling water. Water is
often a scarce resource in sunny regions. The small number of power plants installed
until now have always been able to find local water reserves for the cooling. On a
larger scale freshwater cooling in regions with little water is a problem. In principle,
a solar thermal power plant can also operate with dry cooling. In this case, efficiency

160
7.6 Solar Power Plant Markets

drops a little and the costs increase slightly. On a long-term basis solar thermal
power plants with dry cooling should also become economically interesting enough
to resolve the cooling water issue. If solar thermal plants are installed near the sea,
the water from the sea can provide effective cooling. The waste heat of solar thermal
facilities can also be used to desalinate sea water. It should be possible to produce
carbon-free electric energy and drinking water at the same time but until now this
has not been done on a large scale.

7.6 Solar Power Plant Markets

The biggest markets for power plants exist in countries that have good solar radia-
tion conditions and offer favourable energy credits. The main markets are currently
in Spain and the USA (Figure 7.16).
The main reason for this is the favourable economic conditions in both countries.
Energy credits for solar thermal power plants in Spain are high and above the typical
market prices for electric energy. The USA uses Renewable Portofolio Standards
(RPS). These vary from state to state and establish quotas for renewable energy
plants. Plants are being built or planned in certain US states such as California and
Nevada. Other sunny and hot US states could follow their example.
In addition to Spain, plans are underway to improve the economic conditions for
solar thermal plants in other Mediterranean countries. Italy, Greece and countries
in North Africa could soon be following Spain’s example. In Germany the focus
has been mainly on developing the appropriate technology. Germany is among the
market leaders in the world, specifically in the area of components for solar thermal
parabolic trough power plants. A prototype for a solar tower power plant in Jülich
aims to showcase German technology and attract new export markets.

Figure 7.16 Construction of a parabolic trough collector prototype in Andalusia. Spain is currently one
of the booming markets for solar power plants.

161
7 Solar Power Plants – Even More Energy from the Sun

7.7 Outlook and Development Potential

Although the expansion in solar thermal power plants came to a standstill between
1991 and 2006, many new facilities are now either at the planning stage or under
construction. The current boom in the area of solar power plants is likely to lead to
cost reductions if there is a rapid rise in the power plant capacity installed world-
wide. In about 10 years solar plants at some of the early locations could start becom-
ing competitive with conventional power plants, even without the offer of higher
energy credits. It is conceivable that power plants in the hot and sunny regions of
the world will be fully competitive in about 20 years’ time. Then, purely for cost
reasons, solar power plants will be given preference in these countries over previous
fossil and nuclear power plants.
By far the greatest potential for building solar power plants exists in North Africa.
Even with a generous exclusion of unsuitable areas, such as sand dunes, nature
reserves and mountainous and agricultural regions, about 1% of the remaining area
of North Africa would theoretically be able to produce enough electricity to meet
the demands of every country on earth. Figure 7.17 shows the areas that would be
available. Other perfect solar sites could be found, for example, in the US deserts.

unusable regions:

inclination geomorphology hydrology ocean


land utilisation protectorate population usable

Figure 7.17 Suitable regions in North Africa for building solar power plants. Graphics: DLR.

In addition to the favourable costs of generating energy, solar power plants in North
Africa also have long operating times. With the ability to provide a high number of
full-load hours, the plants there can guarantee a higher reliability of coverage than
those in Central or Northern Europe. North Africa also has many locations with top
conditions for wind power plants.

162
7.7 Outlook and Development Potential

The question is, how can reasonably priced electricity from Egypt or Mauretania
help us to solve our energy problems? The solution to this problem is simple. The
cheap electricity merely has to be transported to us. Figure 7.18 shows the top loca-
tions for obtaining this electricity and the possible transport routes to Europe.

Figure 7.18 Possibilities for importing renewable energy from North Africa to the EU,
along with long-term possible energy generation costs and full-load hours.

Technically as well as financially, the transport of this energy is already viable today
through high-voltage-direct-current transmission. Transmission over a 5000 km-
long cable with losses of less than 15% is possible. These losses would amount to
around 0.5 cents per kilowatt hour, which must be added to the anticipated costs
for electricity generation of 3 to 4 cents per kilowatt hour. Added to this is the cost
of the lines, which works out to between 0.5 and 1 cent per kilowatt hour. Altogether
renewable energy could be produced at 4 to 6 cents per kilowatt hour, transported
to Europe and, in combination with photovoltaic, wind or solar thermal power
plants, guarantee high supply reliability.
At least part of the electricity supply for Central Europe could be covered in the
long term from power plants in Africa – and at a cost similar to our current electric-
ity supply. However, this scenario brings concerns about a new dependency on the
hot countries of North Africa.

163
7 Solar Power Plants – Even More Energy from the Sun

Dependency on one particular state would not arise if power plants were distributed
in a number of locations. Also, importing more than 20 to 30% of the total supply
from one nation would not be wise from the point of view of supply reliability. To
this extent, the risk would be a manageable one. For most of the poor nations of
North Africa the raw material of solar energy could become a positive element in
their economic development and, as a result, contribute towards political stability.
Consequently, both sides would benefit from the import idea.

164
8
8 Wind Power Systems –
Electricity from Thin Air
With all the different renewable energies available today, wind power is considered
a model technology because it has transformed from something of a cottage industry
to a billion dollar business within just a few years. During the mid-1980s wind
power was still viewed as relatively unimportant, yet by 1994 there were already
several thousand installations offering a total output of more than 3500 megawatts.
By 2008, wind wheels turning worldwide were providing an impressive output of
more than 121 000 megawatts. The result was the creation of more than 440 000
new jobs worldwide.
Yet the use of wind power is actually old hat and dates back many centuries. Years
before the birth of Christ simple windmills were being used in the East for irrigation
purposes.
Wind power was not exploited in Europe until much later. In the twelfth century
post mills started to become popular for milling grain in Europe (Figure 8.1). These
had to be rotated towards the wind, either manually or with the help of a donkey.
In addition to milling grain, millers in those days also had the difficult job of man-
aging the mill. When a storm approached, millers had to stop the mill in time so
that they could remove the canvas from the vanes to prevent the power of the wind
from destroying the mill. Wooden block brakes were used to stop the mechanism
but these were not without their own dangers. In their hurry to stop the vanes of the
windmill from turning, millers would sometimes apply the brakes too hard and start
a fire because of the heat that developed on the wooden brake blocks. This is why
so many windmills burned down at the time. Long periods with no wind were
another problem for millers.
In the centuries that followed, mills underwent considerable technical change, as
the Dutch windmill shows. In addition to milling grain, they were now also used to
pump water and carry out other power-driven tasks. Technically sophisticated mills

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

165
8 Wind Power Systems – Electricity from Thin Air

Figure 8.1 Left: Historical post windmill in Stade, Germany. Photo: STADE Tourismus-GmbH. Right: Dutch
windmill on the Danish island of Bornholm.

rotated automatically in the wind and could be braked safely. By the mid-nineteenth
century 200 000 windmills were in use in Europe alone.
By the first half of the twentieth century steam-driven engines and diesel and electric
motors had replaced almost all wind power systems. Wind power did not enjoy a
resurgence until the oil crisis of the late 1970s. More and more sophisticated wind
power plants are now offering a real alternative to fossil and nuclear power
stations.

8.1 Gone with the Wind – Where the Wind


Comes From

The sun is also ultimately responsible for the creation of wind. It bombards our
planet with gigantic amounts of radiation energy, and the earth has to radiate this
incoming solar energy back into space to save itself from continuous warming and
ultimately burning up. However, more solar energy reaches the equator than earth
radiates back into the universe. The path of sunlight to the South and North poles
is longer than to the equator. This explains why the situation at the poles is the exact
opposite. Considerably less solar radiation reaches the poles and more energy is
radiated into space than is emitted by the sun. As a result, a gigantic amount of
energy is transported from the equator to the poles.

166
8.1 Gone with the Wind – Where the Wind Comes From

This transport of heat occurs primarily as a result of a global exchange of air masses.
Major air circulations around the world pump the heat from the equator to the poles.
Gigantic circulation cells, called Hadley cells, develop (Figure 8.2). The earth’s
rotation deflects these currents. This in turn produces relatively stable winds that
for a long time were crucial for sailing ships. In the tropical waters north of the
equator a relatively stable and constant wind blows from the northeast. Unsurprisingly,
it is called the Northeast Trade Wind. In contrast, the Southeast Trade Wind blows
south of the equator.

Figure 8.2 Global circulation and origins of different winds.

Local influences exist in addition to the global wind currents. As a result of the
rotation of the earth, local areas of low and high pressure create wind movements
that rotate around the low pressure point. This rotation runs counterclockwise in the
Northern hemisphere and clockwise in the Southern hemisphere.
The so-called sea-land breeze occurs in areas near the coast. During the daytime the
radiation from the sun warms up the land and the air above it much more than the
ocean water. The warm air above the land rises up and colder air flows in from
the ocean. At night this cycle reverses as land cools off faster than the ocean.
Head winds also occur in the mountains and in polar regions with cold air streaming
down mountain slopes sometimes at very high wind speeds.
Around 2% of solar radiation in the world is converted into wind movement.
Therefore, the energy supplied by wind equates to a multiple of the primary energy

167
8 Wind Power Systems – Electricity from Thin Air

demand of mankind. As is the case with water power, only a very small portion of
this energy is usable. The largest supply of wind power is over the open seas where
no obstacles slow the wind down. Over land the wind very quickly loses its speed
due to the effects of rough terrain. For the same level of wind power as on the open
seas it would be necessary to go to higher altitudes or use substantially larger tracts
of land. Inland the effect on the wind due to the roughness of the terrain is no longer
noticeable at altitudes higher than several hundred metres. As a result, wind use
becomes more difficult as the distance from the coast increases. However, optimal
locations also exist on hills and mountain tops inland.
The supply of wind energy on earth varies considerably (Figure 8.3). Whereas it
could cover about one-third of demand in Germany, the potential in some countries
is so vast that places like Great Britain could even export large quantities of wind
power while covering all their own electric power needs. The best locations for wind
power are usually where the wind hits land from the open seas without slowing down.

Figure 8.3 Average wind


speeds worldwide.
Source: http://visibleearth.
nasa.gov.

8.2 Utilizing Wind

Sailors have been using wind power as driving energy for centuries. However, one
major characteristic of wind power is the strong fluctuation in the energy supplied.
The capacity of the wind increases with the third power of wind velocity. Put more
simply, this means that the capacity of the wind rises eightfold when wind speed
doubles.

168
8.2 Utilizing Wind

Wind Power

For a specific wind velocity v, the power of the wind Pwind is calculated by the area
A for air density ρ:
1
Pwind = ⋅ ρ ⋅ A ⋅ v3
2
With a wind velocity of 2.8 metres per second (m/s), which equates to 10 km/h or
6.2 mph, the wind with an air density of ρ = 1.225 kg/m3 on an area of 1 m2 only
reaches a relatively low capacity of
3
1 kg m
Pwind = ⋅1.225 3 ⋅1 m 2 ⋅ ⎛⎜ 2.8 ⎞⎟ = 13.4 W.
2 m ⎝ s⎠
With a wind velocity of 27.8 m/s, thus 100 km/h or 62.1 mph, the capacity rises a
thousandfold and reaches 13.2 kW, which corresponds to around 18 PS (Figure 8.4).

Figure 8.4 Area through which the wind reaches a capacity of 100 kilowatts at different
wind speeds.

Meteorology still follows the practice of indicating wind velocity v based on the
Beaufort scale (bft). The 12-level Beaufort scale was developed by the British
Admiral Sir Francis Beaufort, who observed the sail behaviour of a naval frigate in
different wind conditions and categorized it into different levels in 1806. The British
navy officially introduced the Beaufort scale in 1838 (Table 8.1).

169
8 Wind Power Systems – Electricity from Thin Air

Table 8.1 The Beaufort wind scale.

bft v in m/s Name Description


0 0–0.2 Calm Smoke rises vertically
1 0.3−1.5 Light air Wind direction only detectable by smoke
2 1.6–3.3 Light breeze Wind noticeable, leaves rustle
3 3.4–5.4 Gentle breeze Leaves and small twigs in constant motion
4 5.5–7.9 Moderate breeze Wind moves branches and thin twigs, raises dust
5 8.0–10.7 Fresh wind Small trees begin to sway
6 10.8–13.8 Strong wind Large branches in motion, overhead wires whistle
7 13.9–17.1 Moderate gale Whole trees in motion, noticeable difficulty in
walking
8 17.2−20.7 Fresh gale Wind breaks branches off trees
9 20.8−24.4 Strong gale Minor structural damage
10 24.5−28.4 Heavy storm Wind uproots trees
11 28.5–32.6 Violent storm Heavy storm damage
12 ≥ 32.7 Hurricane Devastation occurs
10 m/s = 36 km/h = 22.4 mph.

Systems that use wind energy must cope with a strong fluctuation in wind supply.
On the one hand, they have to use the energy supplied by the wind even when wind
speeds are low; on the other hand, they should suffer no damage even when wind
speeds are extreme. Therefore, most wind turbines move into storm mode when
wind velocity is very high.

Wind Speed Records


The strongest gust of wind measured to date had a wind velocity of 412 kilo-
metres per hour (114 metres per second, 256 mph) on 12 April 1934 on Mount
Washington in the USA. The highest ten-minute average of 372 kilometres per hour
(231 mph) was recorded on the same day on Mount Washington. Tornados reach even
higher wind speeds. Radar has established wind speeds of around 500 kilometres per
hour (139 metres per second, 311 mph) with the strongest tornados recorded thus far. At
this rate the wind capacity per square metre totals more than 1.6 megawatts or over 2000
horsepower. This same effect would be achieved if four large lorries of 500 horsepower
each were driven at full speed towards a one square metre surface. On this basis it is not
difficult to understand the extremely destructive effect of tornados.

Modern wind turbines use part of the kinetic energy of the wind. In so doing, they
slow down the wind. In principle, it is not possible to use all the wind’s capacity.
This would require the wind to be stopped completely, and in this case the wind
stream would come to a standstill. This was something the German physicist Albert

170
8.2 Utilizing Wind

Betz recognized when he wrote in 1920 that maximum energy can be extracted from
the wind if it is slowed down to one-third of its original speed. This is the only way
in which theoretically 16/27 or 59.3% of the wind’s potential energy is usable.
Today this theory is called Betz’ Law in honour of the man who made the
discovery.
This power factor therefore identifies how much wind energy a wind turbine uses.
Under ideal operating conditions, optimized modern wind turbines can use just over
50% of the energy contained in wind and convert it into electric energy. They
therefore reach power factors of at least 50%, close to the maximum possible limit.
The formula notation is cp. The power factor is a typical parameter for wind turbines
and is also found in the data sheets of many manufacturers of wind power
generators.
If a wind turbine stops the force of the wind, it ends up changing the course of the
wind’s flow. Behind the wind turbine the braked wind flows through a large area.
The course of the wind’s flow widens (Figure 8.5).

Figure 8.5 A wind turbine


course of flow.

In wind use a distinction is made between


■ Drag principle
■ Lift principle
For example, the old Viking ships that were equipped with square rigs worked on
the drag principle. The sail put up resistance to the wind and the wind pushed the
sail, propelling the ship forward.
With the lift principle, which is also used by modern sailboats with fore-and-aft
sails, considerably more energy can drawn from the wind than with the drag prin-
ciple. Consequently, modern wind turbines function almost exclusively on the basis
of the lift principle. Among the different system concepts, the familiar rotors with
a horizontal axle have been the most successful.

171
8 Wind Power Systems – Electricity from Thin Air

With these turbines, the wind flows from the front towards the rotor hub. Due to
the relatively fast-rotating rotor blades, there is an airstream in addition to the actual
wind and this airstream flows from the side onto the rotor blade (Figure 8.6). The
rotor blade is then struck by a resulting wind that is composed of both the actual
wind and the airstream.

Figure 8.6 Functioning


principle of a wind turbine with
a horizontal axle.

The rotor blade now divides the wind. The resulting wind flows along the rotor
blade. Due to the shape of the rotor blade, the wind has a longer distance to cover
on the upper surface of the blade than on the lower surface. As a result, the flow
widens and the air pressure decreases. Low pressure is created on the upper surface
and high pressure is created on the lower surface. This difference in pressure then
produces a lift force that reacts vertically to the resulting wind.
This lift force can be split into two components: the thrust force in the direction of
the rotor axle and the tangential force in the direction of the circumference.
Unfortunately, in most cases the thrust force is bigger than the tangential force. The
thrust force is not very useful. It merely presses against the rotor blades and bends
them. The construction of the rotor blades themselves must therefore be very stable
so that they can resist the thrust power. Ultimately, the tangential force is responsible
for the rotation of the rotors because it works along the circumference.
Modern wind turbines always try to optimize the volume of tangential force. This
requires an optimal relationship between the actual wind velocity and the airstream.
The rotation of the rotor affects the airstream. The rotation of the rotor blades
enables an optimization of the angle at which the resulting wind approaches the
blade. This deliberate rotation of the rotor blade is known as ‘pitching’.

172
8.3 Installations and Parks

8.3 Installations and Parks

8.3.1 Wind Chargers


One of the applications for smaller wind power systems is charging battery systems.
Technically, an island system with a battery functions in a similar way to a photo-
voltaic island grid, except that a wind generator is used instead of a photovoltaic
module. Small wind power systems are particularly popular for use on boats
(Figure 8.7), where they are often used to charge on-board batteries for anchored
craft.

Figure 8.7 Small wind turbines used to charge battery systems.

Compared to photovoltaic island systems, island grids with wind turbines are techni-
cally more complex in their operation. If a battery is fully charged, the charge regu-
lator of a photovoltaic system removes the photovoltaic module from the battery in
order to prevent overcharging. Photovoltaic modules can cope with this without any
problem. But it is a different case with wind generators that are only separated from
the battery. If a wind generator does not have a load or a battery attached, high wind
speeds can make the rotor blades revolve so fast that the generator is damaged.
Therefore, some wind charge regulators switch wind generators with a fully charged
battery to a heating resistance, thus restricting the number of revolutions.

173
8 Wind Power Systems – Electricity from Thin Air

The electric generators that wind turbines use are almost always alternating current
or three-phase generators. However, a battery can only be charged with direct
current. A rectifier converts the alternating current of the wind power system into
direct current (Figure 8.8).

Figure 8.8 Principle of a simple wind island system.

Small wind turbines are usually mounted onto very low masts. The use of wind is
usually not viable at sites that are far inland or where many trees, houses or other
objects will obstruct its flow. Building regulations should also be checked before-
hand to ensure that a wind system can be erected.
Storms generally pose a threat to wind generators. Large wind turbines therefore
incorporate special storm settings. For cost reasons, small wind turbines often do
not have any built-in storm protection. As a result, it is wise to dismantle the wind
generator as a precautionary measure when large storms are forecast.
In addition to simple wind chargers, more complex island grid systems are available.
These allow wind turbines to be combined with other renewable energy producers.
In many cases, photovoltaic modules and wind generators work relatively well
together. When there is too little sunshine, there is usually a strong wind or vice
versa.

8.3.2 Grid-Connected Wind Turbines


Large wind turbines that feed into the public electricity grid have seen extensive
technical development during recent years. In the 1980s the typical capacity of a
wind system was 100 kilowatts or even less. The size of the rotors was still very
modest at a diameter of less than 20 m (65 ft). In 2005 manufacturers were already
developing prototypes for systems with 5000-kilowatt capacity, in other words 5
megawatts, and rotor diameters of 110 m (360 ft) and more. Today these systems

174
8.3 Installations and Parks

are ready for mass production. Compared to the size of the systems available
today, the traditional windmills and sailing vessels seem like toys (Figure 8.9). They
even dwarf large airliners such as the Boeing 747, which has a wingspan of about
60 m (195 ft), and the mega-airliner A380 with its 80 m (261 ft) wingspan. Apart
from their size, the performance of modern wind turbines is also impressive. For
instance, a single 6-megawatt system can supply all the electricity requirements
for several thousand households in industrialized countries like Germany and
Britain.

5000 kW

1500 kW

500 kW

200 kW
Ø 115 m

90 kW Ø 65 m
Ø 40 m
Ø 30 m
Ø17 m

Figure 8.9 Size development of wind turbines.

However, there are physical limitations to increasing the size of wind turbines. As
the size increases, there is a disproportionate increase in the material requirements.
Furthermore, there are major logistical problems in transporting building parts for
such gigantic structures. Based on current conditions it is therefore highly unlikely
that turbines of over 10 megawatts will be built.
Wind power is currently one of the most effective technologies for combating
climate change. Even the largest wind turbines can be erected in just a few days. A
concrete foundation provides a secure base (Figure 8.10). In places with soft subsoil
posts are used to support the foundation in the ground.
There are three different types of towers that can be used: tubular steel towers, lattice
towers and concrete towers. Years ago lattice towers similar to electricity pylons
were normally used. For aesthetic reasons tubular steel towers then became more
popular. Because of rising steel prices and growing problems in transporting the
large tubular segments needed for the biggest types of turbines, concrete and lattice
towers are now being used more and more. As the size of the turbines grew so too

175
8 Wind Power Systems – Electricity from Thin Air

did the demands placed on crane technology. Today cranes are required to lift many
tons of heavy loads to heights of more than 100 m.

Figure 8.10 Erection of a wind turbine. Top left: Foundation. Right: Tower. Below left: Rotor and
nacelle. Photos: German WindEnergy Association and ABO Wind AG.

The nacelle (Figure 8.11) forms the heart of modern wind turbines. The nacelle is
mounted to the tower in a rotatable state. The wind measurement device establishes
wind velocity and wind direction. The so-called yaw drive then turns the turbine so
that it faces the wind in an optimal position. The individual rotor blades are attached
to the hub. A shaft starts up the movement of the rotor blades and powers the gears
and the electric generator. The purpose of the gears is to adapt the slower rotation
speed of the rotor to the faster rotation speed of the generator. Although the rotor
of a 5 megawatt turbine 126 m in diameter only reaches a rotation speed of a
maximum of 12 revolutions per minutes, the rotor blade tip moves at speeds of over
280 km/h.
Some manufacturers are also promoting wind turbines that do not use gears. Whereas
the electric generator of a wind turbine with gears operates at a rotation speed in
the order of 1000 revolutions per minute, a gearless generator rotates at the rotor
rotation speed. Depending on the size of the wind turbine, this equates to between
6 and 50 revolutions per minute. A generator must be larger in size to use the rela-
tively slower rotor rotation rate. The savings on the gearbox are thus cancelled out

176
8.3 Installations and Parks

wind sensors
Rotor blade

gear
hub
brake

generator
yaw drive

nacelle

tower

Figure 8.11 Structure and components of a wind turbine. Graphics: Nordex AG.

because of a considerably heavier and more expensive generator. However, until


now both wind turbine concepts, that is, with or without a gearbox, have been
equally successful in the turbine market.
Modern wind turbines adapt their rotation speeds to the wind speed. When wind
speeds are low a turbine reduces its rotation speed, thereby making better use of the
wind. Wind speeds of 2.5 to 3.5 m per second, that is, 9 to 13 km/h, are usually
sufficient to start a turbine. Wind turbines reach their maximum capacity with wind
speeds of around 13 m per second (47 km/h or 29 mph). If the wind speed increases
more than that, a turbine begins to limit its speed. It pitches the rotor blades less
favourably into the wind, thereby reducing the momentum. This enables the turbine
to keep its output at a constant level.
When wind speeds are very high at more than 25 to 30 m per second (90 to
108 km/h), a turbine moves into storm mode. The yaw drive turns the entire turbine
out of the wind and brakes hold the rotor firm.
As large wind turbines already produce high output, they usually feed their electric
current directly into a medium-high voltage system. A transformer converts the
generator voltage into the mains voltage.
Maintenance on wind turbines often requires an arduous upward climb for the
technicians, so some turbines are equipped lifts (Figure 8.12). Modern wind turbines
have a relatively low incidence of outages, and a turbine service life of 20 years
can be expected.

177
8 Wind Power Systems – Electricity from Thin Air

Figure 8.12 Maintenance work on wind turbines. Source: REpower Systems AG, Photos: Jan Oelker,
caméléon and Stéphane Cosnard.

8.3.3 Wind Farms


During the pioneering days wind turbines were often erected individually, whereas
today they are almost always grouped in large wind farms (also known as wind
parks). A wind farm consists of at least three turbines but can also have many more.
For example, the Horse Hollow Wind Park erected in 2006 in Texas has 421 wind
turbines with a total capacity of 735 megawatts. These wind turbines are capable
of supplying electricity to 150 000 US households.
The main advantage of wind farms compared to individual turbines is the cost
saving. Planning, erection and maintenance are considerably more economical.
Large wind turbines usually have to have obstacle marking for air traffic. This
includes a coloured mark on the rotor blade tips and navigation lights that come on
when visibility is poor. With wind farms only the outer turbines have a marking.
This saves money and improves the appearance of the farms (Figure 8.13).
The disadvantage of wind farms is mutual interference. If the wind turbines are sited
close together, they can take the wind away from each other. The efficiency of the
turbines at the back is then impacted. It is important that sufficient distance is created
between the turbines in the main direction of the wind so that efficiency losses can
be minimized. However, efficiency loss through mutual interference cannot be
totally prevented. Wind farm efficiency takes into account losses from mutual inter-
ference, which is usually between 85 and 97%. This means that losses between 3
and 15% should be expected.
If wind turbines are too close to housing estates they can have a negative effect due
to noise emission and a shadow being cast as the sun goes down. Turbines should
ideally be erected several hundred metres from the nearest dwellings to avoid creat-
ing this effect.

178
8.3 Installations and Parks

Figure 8.13 Wind parks. Source: REpower Systems AG, Photos: Jan Oelker.

Because of the rules on required distances, it is already becoming increasingly dif-


ficult to find good sites for new wind farms in Germany and Denmark. Therefore,
a greater planning effort is being made in the areas that are still available: the North
Sea and the Baltic.

8.3.4 Offshore Wind Farms


With offshore wind farms the turbines are sited directly in the water. For maximum
effectiveness the depth of the water should not be too great and the turbines should
not be too far from the coast. Along with the enormous size of available area, off-
shore use also provides other benefits: the wind blows stronger and more evenly on
open seas than on land. The power output per wind turbine increases at offshore
sites and can be up to 100% higher than on mainland sites.
Offshore wind turbines differ relatively little from onshore ones. Generally, offshore
turbines do not require a good deal of maintenance. Access to wind turbines on high
seas is not possible in bad weather or when the sea is rough. Special ships are
required for major maintenance work but can only be used if the water is relatively
calm. Because offshore wind turbines stand in salt water, all components have to
be well protected and corrosion-resistant.
Special ships with cranes erect the wind turbines out at sea (Figure 8.14). Special
foundations are used to anchor the wind turbines to the seabed. With monopile
foundations, a large steel pipe is driven many metres into the seabed. The stability
of the bottom is important, and a more secure foundation is needed if the ground is
too soft, which increases the cost. This problem may make it economically unfea-
sible to erect projects in shallow water.
Connecting offshore wind farms to the grid is also more difficult than on land. Sea
cables link the different wind turbines to a transformer station, which looks like a
small drilling platform. The transformer station converts the electric voltage of the

179
8 Wind Power Systems – Electricity from Thin Air

Figure 8.14 The Nysted offshore wind farm in the Baltic Sea off Denmark. Source: http://uk.
nystedhavmoellepark.dk. Construction of the wind farm. Photo: Gunnar Britse.

wind turbines into high voltage to keep transmission losses to a minimum. Direct-
current transmission may also be necessary for installations that are far from shore,
as losses with alternating-current sea cables can be considerable. A special converter
converts the alternating voltage into direct voltage. On land it is converted back into
alternating current. Normal high-voltage lines then transport the electricity to the users.

Offshore, Onshore and Nearshore


The term offshore refers to wind turbines that are sited in the open seas. Use
of the terms onshore and onshore wind farms in respect of wind power is
relatively recent and only become popular in tandem with the growth of offshore wind
parks. Onshore means on land or on the mainland. In this sense onshore wind farms are
all wind farms that are not standing in water. These simply used to be called wind farms.
As part of a test of offshore wind farms, some wind turbines were recently erected in
the water just a few metres from the shore. Technically, these test facilities were really
onshore wind turbines with wet feet. The term nearshore wind turbines has also been
used to differentiate the turbines.

Numerous very large offshore wind farms have been erected in recent years.
Denmark has been one of the pioneers in this field, followed by Britain. By contrast,
Germany has been rather slow in moving ahead in this field. In part this is due to
the major technical requirements involved. German offshore wind farms are sup-
posed to be erected in comparatively deep waters of 20 to 50 m at a distance of 30
to 100 km from the coast (Figure 8.15). The biggest wind farms in Germany would
ideally consist of several hundred turbines with outputs of several gigawatts. The
investment cost would quickly add up to several billion euros.

180
8.3 Installations and Parks

4°0′E 5°0′E 6°0′E 7°0′E 8°0′E 9°0′E

56°0′N 56°0′N

55°0′N 55°0′N

Büsum offshore wind parks


in operation
54°0′N 54°0′N under construction
approved
Cuxhaven planned
not approved

Norden
Wilhelmshaven Bremerhaven grid-connection
Emden approved
Netherlands planned
4°0′E 5°0′E 6°0′E 7°0′E 8°0′E 9°0′E
10°0′E 11°0′E 12°0′E 13°0′E 14°0′E borders
continental shelf
56°0′N 56°0′N
12 nautical mile zone
international border

Sweden water depth


0−10 m
10−20 m
Denmark 20−30 m
30−40 m
40−50 m
50−60 m
55°0′N 55°0′N
mudflat

Flensburg

Schleswig

Kiel Stralsund

Rostock
Greifswald
54°0′N 54°0′N

Wismar
Lübeck

HAMBURG Poland
10°0′E 11°0′E 12°0′E 13°0′E 14°0′E

Figure 8.15 Planning for offshore wind farms in the North Sea and Baltic Sea off Germany.
Source: Federal Maritime and Hydrographic Agency of Germany, www.bsh.de. Status as of January 2008.

181
8 Wind Power Systems – Electricity from Thin Air

From a legal point of view, coastal waters are divided into two areas. Territorial
waters extend to 12 nautical miles (22.2 km) from shore. The exclusive
economic zone then begins and stretches to a maximum of 200 nautical miles
(370.4 km).
Wind farms planned for the exclusive economic zone have to be approved by the
relevant authorities, who must check whether the planned farm will interfere with
shipping or endanger the marine environment. If neither is the case, approval is
granted for a fixed term.

■ www.offshorewind.net North American Offshore Wind Project Information


■ http://www.bsh.de/en German Federal Maritime and Hydrographic Agency

8.4 Planning and Design

Anyone who wants to erect a very small wind turbine in their garden must take into
account local planning laws. Overall heights of up to 10 m are usually not a problem.
It is also important to ensure that no neighbours will suffer from noise or
overshadowing.
Obtaining approval for large wind farms is considerably more difficult. Wind farms
can usually be erected only in specifically designated areas identified within the
regional planning specifications of local authorities and municipalities. Planning and
approval regulations vary depending on the local authority. The local planning
department checks the basic feasibility of a wind farm based on a preliminary plan-
ning application. If everything is in order, the applicant can submit an official
application for the project.
Applications with the relevant forms and explanatory notes can be obtained from
the responsible local authority. In addition to checking the safety of the turbines and
the location, the relevant authority looks into environmental compatibility and the
effect of noise pollution and overshadowing.
Anyone embarking on the sometimes obstacle-strewn path of submitting an applica-
tion should first ensure that the proposed location is technically and economically
suitable for building a wind farm.
The prerequisite for optimal planning is having precise knowledge about the wind
conditions at a chosen site, because even the smallest fluctuation in wind supply
can have a considerable impact on output. If no measurement data on wind velocity
are available for the immediate vicinity of the planned site, it is highly advisable
that a measurement station be set up. This station will chart the wind speeds
over at least one year. These results should then be compared to the long-term
measurement data of other stations and, if necessary, corrected to provide a yearly
average. In the case of commercial wind farms, reliable experts should be

182
8.4 Planning and Design

called upon to carry out measurements and calculations and submit a wind survey
report.
If usable information about the wind speed is available, computer models can be
used to convert the calculations for wind conditions at nearby sites or for other
heights. Data on shadowing caused by other wind turbines can also be incorporated
into these calculations. The calculations should include the frequency distribution
of the wind speed at the pitch level of a wind turbine.

Annual Energy Output of a Wind Turbine

The frequency distribution f(v) indicates how often a wind velocity v occurs during
the year. This usually involves compiling wind velocities at intervals of one metre
per second (m/s). The performance characteristic of a wind turbine indicates the
electric power P(v) a wind turbine produces at wind velocity v. The builder of the
installation usually provides the performance characteristics. The annual energy
output E of a wind turbine is calculated on the basis of both characteristics: the
corresponding value of the frequency distribution for each wind speed is multiplied
by the performance characteristic and then all the results added:
m m m m
E = ∑ f (vi ) ⋅ P (vi ) ⋅ 8760 h = ⎛⎜ f ⎛⎜ 0 ⎞⎟ ⋅ P ⎛⎜ 0 ⎞⎟ + f ⎛⎜ 1 ⎞⎟ ⋅ P ⎛⎜1 ⎞⎟ + …⎞⎟ ⋅ 8760 h
i
⎝ ⎝ s⎠ ⎝ s⎠ ⎝ s⎠ ⎝ s⎠ ⎠
The annual energy output for the characteristics from Figure 8.16 is
E = (0% ⋅ 0 kW + 4.3% ⋅ 0 kW + 8% ⋅ 7.5 kW + 10.8% ⋅ 62.5 kW +
12.3% ⋅ 205 kW + 12.6% ⋅ 435 kW + 11.9% ⋅ 803 kW + 10.5% ⋅1330 kW +
8.6% ⋅ 2038 kW + …) ⋅ 8760 h
= 11 685 000 kWh

13 5500
12 5000
wind power plant power P(v) in kW

11
4500
relative frequency f(v) in %

10
4000
9
8 3500

7 3000

6 2500
5
2000
4
1500
3
1000
2
1 500

0 0
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
0
1
2
3
4
5
6
7
8
9

10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
0
1
2
3
4
5
6
7
8
9

wind power v in m/s wind speed v in m/s

Figure 8.16 Left: Frequency distribution of wind speed. Right: Performance curve of a wind turbine.

183
8 Wind Power Systems – Electricity from Thin Air

With wind farms deductions are made for factors such as mutual interference (farm
efficiency), turbine breakdowns and downtimes (availability) as well as a provision
for other unanticipated losses. Various applications available online offer approxi-
mate calculations of turbine yield.
Professional planning departments use sophisticated computer programmes to
develop even more detailed calculations. However, the principle of the calculations
is always the same as that presented here. In addition to providing calculations on
yield, the object of planning is to provide calculations on wind farm optimization.
Different aspects are considered. Building higher towers and installing larger rotors
will increase the output of an installation. However, this will also increase costs
dramatically. The selection of a certain tower height and rotor size will ultimately
provide the optimum installation from an economic point of view. Furthermore,
losses caused by overshadowing can be minimized through minor changes in the
positioning of individual turbines.

■ http://www.volker-quaschning.de/software/ Online output calculations for


windertrag/index_e.html wind power plants
■ http://www.windpower.org/en/tour/wres/pow

8.5 Economics

Very small wind generators can already be bought for a few hundred euros. A simple
60-watt wind generator for charging batteries costs about 500 euros. This price does
not include the mast and installation. At about 8 euros per watt the costs of wind
generators exceed even those of photovoltaic systems. Therefore, small wind
systems only make economic sense at very windy locations and only then when
electricity does not also have to be fed into a grid.
The costs per watt drop as the wind generator size increases. Around 900 euros per
kilowatt (O/kW), thus around 90 cents per watt, should be included in the calcula-
tions for a grid-connected wind turbine in the megawatt range. The cost of the
turbine, tower and installation is included in this price. The tower and the rotor
blades account for around half of the cost (Figure 8.17).
The ancillary investment costs for planning, development, foundations and grid
connections amount to about 30% of the wind turbine costs (Deutsches Windenergie-
Institut GmbH DEWI, 2002) and need to be added on to the price. Around 1200
euros per kilowatt should be estimated for a turnkey wind farm. The project costs
for a small farm with four turbines of 2.5 megawatts each will amount to 12 million
euros. Depending on the location and the technology, these costs can vary consider-
ably upwards or downwards. A figure of around 5% of the wind turbine costs should
be allocated for annual operation and maintenance.
Wind farms are usually built by operating companies. A group of private investors
extends about 20 to 30% of the cost of erecting the wind farm. The remaining

184
8.5 Economics

installation
3% tower
24 %
wind
tracking
2%
hydraulics
2% rotor blades
24 %
cable and
sensors
3%

hub and
shaf t Figure 8.17 Distribution of
nacelle
6%
8% gear costs for a 1.2-MW wind
generator 18 % turbine (Bundesverband
10 % WindEnergie e.V. BWE/DEWI,
2008).

financing is obtained through bank loans. The feed-in compensation is used for
repayment and redemption costs. The surplus funds are distributed among the inves-
tors. If everything goes according to plan, private investors will receive a return of
6 to 10%. If for whatever reason a wind farm does not achieve the output yield
forecast, the investors can end up losing all the money they put into the project.
Offshore wind farms cost about twice as much as those on land. The prerequisite
for these wind farms is a minimum size of several hundred megawatts. Depending
on the distance from shore and water depth, costs can vary considerably between
different farms.
In Germany the feed-in-tariffs for wind turbines are determined by the Renewable
Energy Law (Table 8.2). These tariffs represent the current costs of wind power
electricity. For onshore wind farms the law defines a basic compensation and a
higher rate of compensation. Compensation for feed-in electricity is granted over a
20-year period. Wind parks at very favourable locations on land receive the higher
rate over five years; at less optimal sites this compensation is extended over the

Table 8.2 Compensation in cents/kWh for wind turbines in Germany, based on the Renewable
Energy Law.

In operation 2004 2005 2006 2007 2008 2009 2010 2011


Onshore higher-rate 8.70 8.53 8.36 8.19 8.03 9.20 9.11 9.02
compensation
Onshore basic 5.50 5.39 5.28 5.17 5.07 5.02 4.97 4.92
compensation
Offshore higher-rate 9.10 9.10 9.10 9.10 8.92 15.00 15.00 15.00
compensation
Offshore basic 6.19 6.19 6.19 6.19 6.07 3.50 3.50 3.50
compensation

185
8 Wind Power Systems – Electricity from Thin Air

entire 20 years. On the other hand, erecting wind turbines at poor sites is not encour-
aged. With offshore installations the compensation structure is similar, except that
other rates of compensation do exist and the higher rate remains in effect for at least
12 years. The compensation increases at a minimal rate if wind turbines make a
technical contribution to grid stability or when old wind turbines are replaced by
new ones (repowering).
The profitability of a site is gauged by the number of full-load hours. A good inland
site achieves about 2000 full-load hours per year. In other words, a wind turbine
generates as much electricity at the site during the course of a year as if it were
operating 2000 hours at full load in one continuous segment of time.
Wind turbines at poor inland sites reach noticeably fewer than 2000 full-load hours.
It is easier to improve values in mountainous areas and at the coast. Offshore wind
turbines can reach 3000 full-load hours or even more.
Figure 8.18 shows electricity generation costs for wind turbines based on different
full-load hours. The assumption for all calculations is that the yearly operating and
maintenance costs amount to 5% of the investment costs.

0.22 0.22
800 800
1000 1000
0.20 1200 1200 0.20
electricity production costs in €/kWh_

1400 1400

electricity production costs in €/kWh_


0.18 1600 1600 0.18
1800 1800
2000 2000
0.16 2200 2200 0.16
2400 2400
0.14 2600 2600 0.14
2800 2800
0.12 €/kW €/kW 0.12

0.10 0.10

0.08 0.08

0.06 0.06

0.04 0.04

0.02 0.02
return 0 % return 6 %
0.00 0.00
1200 1600 2000 2400 2800 3200 3600 4000 1200 1600 2000 2400 2800 3200 3600 4000
full-load hours in h/a full-load hours in h/a

Figure 8.18 Electricity generation costs for wind turbines based on full-load hours and investment
costs in O/kW. Assumption: Annual operating and maintenance cost proportion is 5% of the investment cost.

Consequently, on the basis of an inland site reaching 2000 full-load hours and with
investment costs of 1200 euros per kilowatt (O/kW), the generating costs amount to
8 cents per kilowatt hour (0.08 O/kWh) – resulting in an overall return of 6%. This
is the minimum return that should be accepted to cover the financial risks involved.
For an offshore wind farm with 3400 full-load hours per year and investment costs
of 2800 O/kW, the generating costs are over 0.11 O/kWh. As the risk involved for
the first offshore wind farms is considerably higher than for those on mainland sites,
the cost must be offset by a higher return. This will affect the economic feasibility
of a project, which explains the hesitation to develop offshore wind farms before

186
8.6 Ecology

2008. A reduction in the cost of energy generation is expected in future as invest-


ment and operating costs decrease and system yield is optimized.

8.6 Ecology

Because of their size, wind turbines are often visible even from great distances and
therefore can create an eyesore in many places. Whether one finds wind turbines
attractive or ugly is really a matter of personal taste. Nevertheless, the subject has
become a major bone of contention. Whereas supporters swoon over the majestic
appearance of these technical masterpieces, opponents fight against what they con-
sider a blot on the landscape. There certainly are arguments that can legitimately
be made against wind turbines. However, it is important to bear in mind that wind
turbines are usually erected in areas that already bear many signs of human activity.
Traditional windmills have been a common sight for many centuries, and are now
regarded as picturesque additions to the landscape. Domestic livestock seems to get
used to the turbines much more quickly than humans (Figure 8.19). It is incredible
to think that around 190 000 high-voltage pylons and more than one million medium
and low-voltage masts are less controversial in Germany than the 20 000 wind tur-
bines there. The strong feelings about wind turbines may arise because people are
less used to seeing them than the other power infrastructure that criss-crosses the
European landscape.

Figure 8.19 Wind turbines in a rural landscape. Source left: German WindEnergy Association, Source right:
REpower Systems AG, Photo: Jan Oelker.

Considered objectively, the negative effects can be kept to a minimum. Erecting wind
turbines in conservation areas is just as much a taboo as it would be in residential
areas. If wind turbines adhere to a minimum distance from residential areas, the

187
8 Wind Power Systems – Electricity from Thin Air

nuisance from noise or overshadowing will be very minor. These issues are part of
the approval process in the planning of wind farms.
Wind power has very little effect on the animal world. Most animals get used to the
turbines quickly. Birds usually detect the relatively slow-rotating rotors from a great
distance and fly around them. Despite the claims made by the opponents of wind
farms, birds very seldom fly into a turbine. The numerous window panes in resi-
dential buildings pose a far greater threat to the bird world than wind turbines.
The effect that wind power has on protecting the environment is very positive. In
comparison to photovoltaics, the energy required for producing wind turbines is
minimal. A wind turbine will recoup this amount within a few months. Depending
on the site, it would save up to 1 kg of carbon dioxide per feed-in kilowatt hour.

8.7 Wind Power Markets

Denmark is rightfully considered the pioneer of modern wind power. The Danish
teacher Paul la Cour built a wind turbine for generating electricity at the primary
school in Askow as far back as 1891. When the current wind power boom began
in the 1980s, Danish companies ranked among the leaders of this technology from
the start. Today the Danish Vestas Group is one of the most successful wind energy
companies in the world, employing around 20 800 staff worldwide in 2008 and
achieving a turnover of 6 billion euros.

130 GW
120 others
China
110
Denmark
100 India
90 Spain
USA
80 Germany
70
60
50
40
30
20
10
0
1994

1995

1996

1997

1998

1999

2000

2001

2002

2003

2004

2005

2006

2007

2008

Figure 8.20 Development of wind power capacity installed worldwide.

The USA experienced its first wind power boom at the end of the 1980s, by which
time it had already erected thousands of wind turbines in just a few years. However,
the wind power market in the USA came to an almost complete halt in the early
1990s. Due to a legally regulated system of feed-in-tariffs, Germany then became

188
8.8 Outlook and Development Potential

number one in the wind energy market in 1990s. As a result, Germany developed
an internationally successful wind energy industry, with a turnover of 7.6 billion
euros in 2007 and an export quota of over 83%.
Whereas the number of new wind farms in Denmark has dropped to almost zero in
recent years, expansion in Germany is continuing. Germany, the USA, China and
Spain are the key wind energy markets today. In Spain a legal system of feed-in-
tariffs similar to that in Germany is ensuring continued expansion. The wind energy
market in the USA is driven by tax write-offs and quotas for renewable energies.
Due to changing conditions in the past few years, the market there has been very
up and down.
India has also developed a steady wind energy market and, consequently, a flourish-
ing wind energy industry. The Indian Suzlon Group is one of the biggest wind
turbine manufacturers in the world. Numerous other countries have started to
develop wind energy use during the last few years, and, as a result, high annual
growth rates are anticipated in the wind energy sector (Figure 8.20).

■ www.awea.org American Wind Energy Association


■ www.ewea.org European Wind Energy Association
■ www.gwec.net Global Wind Energy Council

8.8 Outlook and Development Potential

Whereas only very few countries were players in the wind boom of the 1990s, many
more have started to rely on wind energy in the meantime. Even today at good sites
wind energy is able to compete with conventional fossil power plants. In contrast
with the fluctuating prices for coal, natural gas and crude oil, the production costs
for a wind turbine once it has been erected are quite constant. It is therefore antici-
pated that the high growth rates for wind power will be sustained worldwide.
Germany can be viewed as a prime example of a country that has fully exploited
the potential of its wind energy. On the other hand, its market has been shrinking
since 2002 (Figure 8.21). The reason is that the best inland sites have already been
developed. Presumably wind turbines will be standing on all economically feasible
and legally viable sites by 2015. As wind turbines have a service life of about 20
years, Germany is successively developing a new market segment: repowering.
With repowering, new more efficient installations replace older and no longer viable
smaller wind turbines. As a result, the installable wind energy output is increased
on land. The total potential is a minimum of 30 000 megawatts. These installations
could produce around 60 billion kilowatt hours per year, which equates to about
10% of current electricity demand.
Germany is beginning to develop offshore sites, a process that is expected to reach
its peak in about 2020. Around 30 000 megawatts could be installed in the offshore
areas by 2030. Due to the higher wind supply offshore, these installations would

189
8 Wind Power Systems – Electricity from Thin Air

6000
Offshore Repowering
MW
Offshore
5000
Onshore Repowering
Onshore
4000

3000

2000

1000

0
1990

1992

1994

1996

1998

2000

2002

2004

2006

2008

2010

2012

2014

2016

2018

2020

2022

2024

2026

2028

2030
Figure 8.21 Development of newly installed wind power capacity in Germany and
projects until 2030. (Data based on: Messe Hamburg Wind Energy, 2006)

produce around 100 billion kilowatt hours per year. Overall wind energy could cover
between 25 and 30% of Germany’s electricity requirements by 2030.
According to the American Wind Energy Association (AWEA) about 20% of the
US American electricity demand can be covered by wind power by 2030. To reach
this goal the annual installations of new wind turbines in the USA would need to
ramp up to more than 16 gigawatts by 2018 compared to 8.5 gigawatts in 2008.
That would lead to a cumulative total power of 300 gigawatts. In contrast to
Germany, more than 80% of the installation will be land-based.
The possible installations for the European Union are in the same range. The
European Wind Energy Association (EWEA) projects that the total installed
capacity by 2030, will be between 200 and 350 gigawatts. These figures provide
convincing evidence that wind power will become one of the key pillars of a
climate-compatible electricity economy. As other countries are in the process of
following the US and Europe’s model, wind power could very possibly provide 25
to 35% of the world’s electricity needs by 2050.

190
9
9 Hydropower Plants –
Wet Energy
In terms of numbers, there are far fewer water-powered systems today than there
were at their heyday at the end of the eighteenth century. At that time between
500 000 and 600 000 watermills were turning in Europe alone (König, 1999). In
those days France was the country that had the highest number, although thousands
were also being used in other parts of Europe. Water wheels not only drive mills
but also power a number of other work and tool machines. Flowing waterfalls were
harnessed by watermills with turning wheels up to 18 m in diameter. At five to seven
horsepower the average power of water wheels at that time was still comparatively
modest (Figure 9.1).

Figure 9.1 Historic watermill in


the Alps. Source: www.verbund.at.

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

191
9 Hydropower Plants – Wet Energy

As more and more mills were built on rivers and streams, the activity was strictly
regulated and mill operators were instructed on how long mills could be used and
how large they could be. This may have been an annoyance for them, but it was a
good thing because it promoted technical development and ensured that optimal use
was made of existing mills. This constant drive to improve gave rise to the modern,
highly efficient turbines used in hydropower plants today.
The introduction of the steam engine slowly displaced water-powered systems. But,
in contrast to wind power, the use of hydropower did not vanish from the scene
with the exploitation of fossil energies. When widespread electrification began at
the end of the nineteenth century, hydropower was still very much part of the scene.
At the beginning small turbines were used to power electric generators, but the size
of the systems grew rapidly.

9.1 Tapping into the Water Cycle

The colour of our planet is blue when seen from space. The reason is that 71% of
the earth’s surface consists of water. However, without the sun our blue planet
would not be blue. Water, which gives the earth its characteristic appearance, would
be completely solidified into ice. Because of the heat of the sun, 98% of water is
fluid.

A Water-Powered System in the Rain Gutter?


The roof of a house collects many cubic metres of water each year. A rain
gutter diverts the water without making any use of its energy. A water-
powered system for each rain gutter could actually be a good idea.
The annual precipitation in Berlin amounts to around 600 litres per square metre, so that
a 100 m2 house roof would collect 60 000 litres or 6000 10-litre buckets of water.
Compared with a similar bucket on the ground, the content of a 10-litre bucket on a
10 m-high roof has a potential energy of 0.000273 kilowatt hours. Six thousand buckets
together create around 1.6 kilowatt hours, just enough to boil water for 80 cups of coffee
but unfortunately too little to make it worthwhile to harness. Thus considerably larger
quantities of water than are available in a rain gutter are necessary to produce viable
quantities of energy.

Altogether there are around 1.4 billion cubic kilometres of water on earth. Of this
amount, 97.5% is salt water in the oceans and only 2.6% is fresh water. Almost
three-quarters of the fresh water is bound in polar ice, ocean ice and glaciers; the
rest is mainly in the groundwater and the soil moisture. Only 0.02% of the water
on earth is in rivers and lakes.
Due to the influence of the sun, on average 980 litres of water evaporate from each
square metre of the earth’s surface and come down again somewhere else in the

192
9.1 Tapping into the Water Cycle

form of precipitation. Altogether about 500 000 cubic kilometres of water collect in
this way each year. This gigantic water cycle converts around 22% of all the solar
energy radiated onto earth (Figure 9.2).

Figure 9.2 Earth’s water cycle.

If the evaporation were concentrated onto a single square kilometre on earth, the
water would come pouring out at a speed of over 50 km/h. The column of water
would reach the moon in less than one year. Fortunately, however, the water remains
spread on the surface of the earth, because otherwise we would not have any liquid
water left in 3000 years.
The energy that is converted by the sun in the earth’s water cycle is around 3000
times the primary energy demand on earth. If we were to stretch a tarpaulin at a
height of 100 m around the earth, collect all the rainwater into it and use it to produce
energy, we would be able to satisfy all the energy needs on earth.
Altogether 80% of precipitation comes down over the ocean. Of the 20% that falls
on land, a large amount evaporates. Around 44 000 cubic kilometres of water
reach the ocean again as return flow in groundwater or in rivers. This is still
more than one billion litres per second. We could be making use of the energy of
the water from this return flow without stretching large tarpaulins over the earth.
This water could be delivering some of its energy on its way back to the ocean –
energy that it first received from the sun. Yet the useable amount of energy that
rivers carry with them comprises only a small fraction of the energy of the water
cycle.
To convert water into power, it is not just the amount of water that is important
but the height at which water is found. Water in a small mountain stream with

193
9 Hydropower Plants – Wet Energy

many hundreds of metres of difference in elevation can sometimes carry more


energy than a large river that just has to descend from an elevation of a few
metres to reach the ocean. About a quarter of the energy of water from rivers and
lakes can theoretically be used. This equates to around 10% of the current primary
energy requirement worldwide. Ocean currents and waves also contain useable
energy.

9.2 Water Turbines

Water turbines form the core of water-powered systems and extract the energy from
water. Modern water turbines have very little in common with the rotating wheels
of traditional watermills. Depending on the head of the water and the water flow,
turbines optimized for the respective operating area are used (Figure 9.3). These
turbines reach a power of over 700 megawatts.

Figure 9.3 Operating areas for different water-powered turbines.

The Kaplan turbine, developed by the Austrian engineer Viktor Kaplan in 1912, is
usually the first choice for low heads – for instance, for power plants on rivers
(Figure 9.4). This turbine uses the pressure of the water at the different elevations
of barrages. It has three to eight adjustable blades and looks like a large ship screw
(propeller) that powers the flowing water. The efficiency of Kaplan turbines reaches
values between 80 and 95%.
The bulb turbine (Figure 9.5) is similar to the Kaplan turbine, except that it has a
horizontal axle and, as a result, is suitable for even smaller heads. The generator is

194
9.2 Water Turbines

Figure 9.4 Drawing showing a Kaplan turbine with a generator (left) and a photo of a Kaplan turbine
(right). Source: Voith Hydro.

placed in a bulb-shaped workroom behind the turbine, which explains why it is also
referred to as a bulb turbine.

Figure 9.5 Bulb turbine with


generator. Source: Voith Hydro.

The Francis turbine, which was developed by the Briton James Bicheno Francis in
1848, is used for larger heads of up to 700 m. This turbine also uses the pressure
difference of water and reaches efficiencies of over 90%. In principle, the Francis
turbine can also be used as a pump and is therefore suitable as a pump turbine for
pumped-storage plants (Figure 9.6).

195
9 Hydropower Plants – Wet Energy

Figure 9.6 Francis pump turbine at Goldisthal pumped-storage plant (left) and a Francis turbine at the
Itaipu plant (right). Source: Voith Hydro.

In 1880 the American Lester Allen Pelton developed the Pelton turbine. It is mainly
designed for large heads and, consequently, for use in high mountains. This turbine
can reach very high efficiencies of 90 to 95%. The water is supplied to the turbine
over a penstock. The water then flows through a nozzle at very high velocity to
spoon-shaped buckets (Figure 9.7).

Figure 9.7 Drawing of a 6-nozzle Pelton turbine (left) and photo of a Pelton turbine (right). Source: Voith
Hydro.

196
9.3 Hydropower Plants

Small plants use the Ossberger turbine, which is also called a cross-flow turbine.
These turbines reach somewhat lower efficiencies of around 80%. The turbine is
divided into three parts that can be powered separately by water. This helps the
system cope with fluctuations in the water runoff of small rivers.

9.3 Hydropower Plants

The energy that can be exploited from water essentially depends on two parameters:
runoff volume and the head of the water. Almost all hydropower plants utilize
natural elevation differences using technical equipment.

9.3.1 Run-of-River Power Plants


The natural course of a river itself concentrates large quantities of water. A run-of-
river power plant can be built anywhere on a river where a sufficient difference in
elevation exists (Figure 9.8). The weir then backs up the water. This creates a dif-
ference in elevation in the water’s surface above and below the plant. At the top
the water flows through a turbine that powers an electric generator. Grating at the
turbine intake prevents rubbish and flotsam washed along by the river from blocking
up the turbine. A transformer then converts the voltage of the generator into the
desired mains voltage.

Figure 9.8 Principle of a run-of-river power plant.

Large hydropower plants are usually constructed so that multiple turbines can run
in parallel. If the water flow drops during the dry periods of the year, some of the
turbines can be shut down. The remaining turbines then still receive almost the full
amount of water they need. This prevents the turbines from working in partial-load
mode with poor efficiency. If, on the other hand, there is flooding and a river carries
more water with it than the turbines can process, the surplus water has to be let out
unused over a weir.

197
9 Hydropower Plants – Wet Energy

Barrages can present an obstacle for shipping and other forms of life on a river.
Locks installed parallel to a barrage enable ships or boats to overcome the difference
in water height. A certain amount of residual water flows through a fish bypass
system next to the weir. This enables fish and other inhabitants of the water to move
freely across the weir system.
Many plants, such as the run-of-river plant at Laufenberg on the Rhine (Figure 9.9),
were built in the first half of the twentieth century. Existing plants that have been
modernized have been able to improve performance. The best example in Germany
is the Rheinfelden power plant, which was built in 1898 and is currently being
modernized. The new facility, which is due to be completed in 2011, will increase
power from 25.7 to 100 megawatts and more than triple electricity production. At
the same time modern fish bypasses are improving the ecological situation.

Figure 9.9 Run-of-river power plant in Laufenburg, Germany. Photo: Energiedienst AG.

Europe has very strict environmental conditions for new energy plants and for those
being modernized. Consequently, it is not anticipated that electricity generation
from run-of-river power plants will increase substantially.
As the height difference in water levels with run-of-river plants is usually only a
few metres, very few plants have an output of more than 100 megawatts of electric
power. Run-of-river plants are also usually difficult to regulate. They supply elec-
tricity around the clock, but because the currents of a river cannot be controlled,
surplus water remains unused if the power is not needed.

9.3.2 Storage Power Plants


Storage power plants produce high levels of power output. Dams store huge masses
of water at geographically optimal locations. This type of dam makes it possible for

198
9.3 Hydropower Plants

storage power plants to be built in the mountains (Figure 9.10). A high-pressure


pipeline pumps the water into a machine house, where, due to the large head, enor-
mous water pressure of up to 200 bars is created. In the machine house water powers
the turbines that produce energy over an electric generator.

Figure 9.10 Storage power plants with reservoirs: Malta (left), Kaprun in Austria (right). Source:
www.verbund.at.

It is not unusual to find dams over 100 m high. The highest dams on earth are over
300 m high. Reservoirs are often also used to store drinking water and to control
flooding.
Storage power plants that are designed primarily to generate electricity have very
high power output. Power plants that produce several hundred or even thousand
megawatts are not unusual (Table 9.1).

Table 9.1 The largest hydroelectric power plants in the world.


Power plant Country River Completed Power Dam Dam
output length height
in MW in m in m
Three Gorges China Yangtse 2009 18 200 2310 180
Dam
Itaipú Brazil/ Paraná 1983 14 000 7760 196
Paraguay
Guri Venezuela Rio Caroni 1986 10 300 1300 162
Tucuruí Brazil Rio Tocantins 1984 7 960 6900 78
Grand Coulee USA Columbia River 1942 6 495 1592 168

199
9 Hydropower Plants – Wet Energy

Storage power plants can be constructed with a second basin located at a lower
elevation at the machine house. Special pump turbines can then pump the water
from the lower basin back up to the higher basin. This is called a pumped-storage
power plant.

9.3.3 Pumped-Storage Power Plants


Pumped-storage power plants need very specific geographical conditions. To be
built, they need two basins with a major difference in altitude between them. Some
pumped-storage plants have a natural inflow where a river flows into the higher
basin. Pumped-storage plants without a natural inflow are purely storage plants.
When electric power is needed, the water of the higher basin is pumped over an
intake mechanism through a pressure pipe to the turbine, which drives an electric
machine as the generator. Once the turbine has extracted the energy from the water,
the water flows into the lower basin. A transformer there converts the voltage of
the generator to mains voltage (Figure 9.11).

Figure 9.11 Principle of a pumped-storage power plant.

When electricity surpluses occur in the grid, the pumped-storage plant switches over
to pumping operation. The electric machine then functions as a motor that drives
the pump turbine. The pump turbine pumps the water from the lower to the higher
basin once again. Major fluctuations in pressure can occur when the generator is
switched to pumping operation, and, in an extreme case, these fluctuations can cause
damage to the pressure pipe or other system parts. A surge tank regulates this change
in pressure.
Pumped-storage power plants reach efficiencies of 70 to 90%. A good 70% of
the electric energy that is necessary for pumping the water to the higher level can
be replaced during the operation of the generator. Despite the losses, pumped-
storage plants are very attractive economically. When a surplus of supply exists,
electricity can usually be obtained very inexpensively. On the other hand, if
electricity is in low supply, a plant can feed the energy back into the grid at a higher
price.

200
9.3 Hydropower Plants

Pumped-storage power plants have become much more common, particularly in


recent years. Supply fluctuations frequently occur in the generation of wind power
in particular. Large pumped-storage power plants can at least partially compensate
for these fluctuations, thereby helping to improve integration of wind power into
the grid.
The four generators of the Goldisthal power plant (Figure 9.12), which was com-
missioned in 2003, together have an output of 1060 megawatts. The accumulated
dam volume of the upper basin is 12 million cubic metres. Based on this volume
and with a medium head of 302 m, the power plant can work at full capacity for
eight hours. This is sufficient to cover the power demands of over 2.7 million
average households.

Figure 9.12 The Goldisthal pumped-storage power plant in Germany. Source: Vattenfall
Europe.

9.3.4 Tidal Power Plants


Tidal waves are attributed to the interaction of the forces of attraction between
moon, sun and earth. As a result of the rotation of the earth, the forces of attraction
are continuously changing their direction. The water masses of the oceans follow
the attraction. As a result, a tidal wave with a height difference of more than one
metre can form on the open sea. Tidal waves caused by the moon occur approxi-
mately every 12 hours at some point on earth. In extreme cases, they can reach more
than 10 m in height. Tidal power plants could use these changes in water level to
generate energy.
In regions with a high tidal hub, a reservoir can be built so that water flows into
it at high tide. The incoming water flows into the reservoir through a turbine in
the dam wall, and at low tide it flows back again. The turbine and the connected

201
9 Hydropower Plants – Wet Energy

generator convert the energy of the water into electric energy. The energy output is
not continuous, so that when the tide goes out output drops to zero.
Tides were used by tidal mills even during the Middle Ages. Worldwide there are
only very few modern tidal power plants in operation today. The biggest and best-
known one is the Rance power plant in France, which is situated on the estuary of
the river of the same name and came into operation in 1967. It has a power output
of 240 megawatts. The dam wall that seals off the 22 km2 basin is 750 m long. The
environmental impacts of the separation of the estuary from the ocean are consider-
able. Corrosion by salty sea water is also causing some problems.
Tidal power plants are comparatively expensive to operate, so it is not likely that
many new plants will be built. Due to the relatively minor differences between ebb
and flow, most countries do not have suitable locations for these plants.

9.3.5 Wave Power Plants


For decades high hopes have been pinned on the development of wave power plants.
A study of the potential of wave energy indicates that large amounts of energy could
be generated. However, only coastal regions with low water depths are appropriate
for the use of wave energy. Due to the comparatively small useable sea areas avail-
able in many countries, this technology has relatively low potential.
Basically, a distinction is made between the following functions:
■ Float ball systems
■ Chamber systems
■ Tapchan systems
Float ball systems use the energy potential of waves. A float ball follows the move-
ment of waves. Part of the installation is anchored to the ground. The movement of
the float ball can be used by a piston or a turbine.
With chamber systems, a chamber locks in air. The waves cause the water level to
fluctuate in the chamber. The oscillating water level compresses the air. The dis-
placed air escapes through an opening and powers a turbine and a generator. When
the water column falls, the air flows back through the turbine into the chamber
(Figure 9.13).
‘Tapchan’ is an abbreviation for ‘tapered channel’. In these systems, waves in
coastal areas or on a floating object feed into a tapered and rising channel. An upper
basin captures the waves. When the water flows back into the ocean, it powers a
turbine.
Although numerous prototypes have been developed for wave power plants in recent
decades, they have not yet caught on in a major way. The main problem is the
extreme fluctuation in conditions at sea. On one hand, technical systems must be
designed to save on materials and be cost-effective. On the other hand, storms with
giant waves place extreme pressure on system survival. Many prototype installations
have already fallen victim to storms. However, large companies have now become
involved in the development of wave power plants, so it is likely that these problems
will be solved in due course.

202
9.3 Hydropower Plants

Figure 9.13 Principle of wave power plants. Left: Float ball system. Right: Chamber system.

9.3.6 Ocean Current Power Plants


Ocean current power plants have a structure similar to wind power plants, except
that the rotors rotate below the water surface. A built-in hub mechanism raises the
rotor to the water surface for maintenance. The first prototype was successfully
installed in 2003 off the North Devon coast in England (Bard et al., 2004)
(Figure 9.14).

Figure 9.14 Left: Prototype system in Project Seaflow off the west coast of England. Photo: ISET.
Right: Maintenance ship in a planned ocean current power plant park. Graphics: MCT.

203
9 Hydropower Plants – Wet Energy

In principle, the physical characteristics of wind power plants are similar to those
of ocean current plants. The main difference is that, because water has a higher
density than air, ocean current plants can achieve higher output yields than wind
power plants even when the speed of currents is low.
Ocean current plants are limited to regions with relatively consistent high current
speeds and moderate water depths of up to 25 m. These conditions mainly exist on
headlands and bays, between islands and in straits. Although shipping lanes often
restrict this kind of use, major potential exists. As technological advances and the
benefits of mass production are rapidly leading to cost reductions, in the medium
term ocean current power plants could become another building block in the genera-
tion of a climate-compatible electricity supply.

9.4 Planning and Design

All types of hydropower plants require different and sometimes complex planning
and design. It is only possible here to outline some of the basic planning aspects.
The first requirement for a run-of-river power plant is the collection of information
about the stretch of water where the plant is to be built. The most important param-
eter is the runoff of the river’s water over the course of a year – that is, the volume
of water that flows through the river. Each river has its own typical annual flow
characteristic that is influenced mainly by rain and melting snow. For further design
purposes, the river runoff volumes are sorted to obtain an annual continuous curve.
This indicates the number of days per year on which a river achieves or exceeds a
certain runoff volume (Figure 9.15).

2500 2500
duration curve of runoff in m³/s

2000 2000
runoff over a year in m³/s

1500 1500

not
1000 usable 1000
rated runof f

part load
500 500
usable
water

0 0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
0 30 60 90 120 150 180 210 240 270 300 330 360
days per year

Figure 9.15 Annual flow characteristics and annual continuous curve for the Rhine runoff near
Rheinfelden.

204
9.4 Planning and Design

Then a rated runoff is established. This is the quantity of water a power plant needs
to produce full output. If the runoff volume of the river rises above the rated runoff,
the surplus water must be diverted unused over a weir. If a large quantity of elec-
tricity is to be generated, turbines with a high rated runoff should be selected.
However, if the runoff volume of the river drops below the rated runoff, there will
not be enough water to enable the power plant to provide its full output. The turbines
will then either work at poor efficiency in partial load or individual turbines will be
switched off and remain unused. A low rated runoff should be selected to ensure
that optimal use is being made of the turbines. In practice, the rated runoff is deter-
mined based on a compromise between the maximum amount of power to be gener-
ated and optimal use of the turbine.

Power Output of a Run-of-River Power Plant

If the water runoff quantity Q and the head H of the water at a power plant are
known, then the electric power output of a hydropower plant can be calculated rela-
tively easily based on the efficiency η of the plant, the density of the water ρW
(ρW ≈ 1000 kg/m3) and the gravitational acceleration g (g = 9.81 m/s2):
Pel = η ⋅ ρW ⋅ g ⋅ Q ⋅ H

With an extension runoff quantity of 1355 m3/s, a head of 10.1 m and efficiency of
79% = 0.79, the output for full load at the Laufenburg hydropower plant on the
Rhine can be calculated as
kg m m3
Pel = 0.79 ⋅1000 ⋅ 9.81 ⋅ 1355 ⋅10.1 m = 106 MW.
m3 s2 s
At 5940 full-load hours per year, the power plant generates 106 MW · 5940 h =
630 000 MWh = 630 million kWh. This is sufficient to cover the average electricity
consumption of 180 000 households in Germany.

Approval granting water rights is required to build a hydropower plant. Normally


approval is granted for 30 years, but it may be given for a longer time. The procedure
for obtaining approval can take three to ten years. For large-scale power plants the
approval also includes a test for environmental compatibility. The EU water rights
guidelines stipulate that all water bodies in Europe must be kept in a ‘good state’.
Although the water may be used, its ecological functioning must remain intact. In
this context the current state of the water body concerned is also crucial. Therefore,
it is much easier to gain approval to modernize a power plant on a stretch of water
that has already been greatly altered than to build a new plant. Because of this, the
chances of obtaining approval to build new large-scale hydropower plants are not
very favourable.

205
9 Hydropower Plants – Wet Energy

9.5 Economics

Hydroelectric power plants are considered the most cost-effective option for sup-
plying renewable electricity today. This applies mainly to older systems where the
construction costs have largely already been amortized. The relatively high con-
struction costs and long amortization periods for new sites substantially increase the
cost of generating electricity.
For small plants fewer than five megawatts the investment in modernization amounts
to between 2500 and 5000 euros per kilowatt; for the reactivation of a plant or a
new plant it is between 5000 and 13 000 euros per kilowatt. The costs for larger
plants are somewhat lower but depend greatly on local conditions. In addition to
the investment and operating costs, large-scale hydropower plants have to pay a fee
for water use in some countries.
With old medium-sized plants in the power range between 10 and 100 megawatts,
the cost to generate electricity is less than 2 cents per kilowatt hour. For new plants
this cost can increase up to 4 to 10 cents per kilowatt hour (Fichtner, 2003). The
costs for small plants can even be higher.

9.6 Ecology

Hydropower plants are the most controversial of all the renewable power plants.
Classic run-of-river, storage and tidal power plants in particular have a huge impact
on the natural environment.
Due to the barrage system, bodies of stationary water form in places where water
flowing over pebbles and boulders previously provided a suitable habitat for many
different kinds of fish. As a result of the changes in their habitat, numerous fish and
plants are becoming extinct. Another danger for fish is the hydroelectric turbines
themselves. Although grating prevents large fish from entering the processing
area of a plant, the metal bars of the grating let small forms of life slip through
and become injured or killed by the turbines. The barrier systems are often an
insurmountable obstacle for fish swimming in the water. Fish bypasses that run
alongside the barriers help enormously in improving the freedom of movement for
fish (Figure 9.16). Nevertheless, barrier systems are still an obstacle for certain types
of fish.
Large reservoirs flood wide areas of land and destroy people’s homes as well as
natural habitat. Sinking biomass decomposes in water and releases large quantities
of climate-damaging methane. This problem can be reduced considerably if reser-
voir basins are carefully cleared before they are flooded.
Breaches in walls are another danger with large dams. Normally dams are con-
structed so that they are largely earthquake-proof. But even the best construction is
powerless against targeted terrorist attacks. Huge devastation will occur if the stored
up masses of water pour into a valley all at once.

206
9.7 Hydropower Markets

Figure 9.16 Left: Stream in the area of the Donaukraftwerk Freudenau power plant in Austria. Source:
www.verbund.at. Fish bypass at Weserkraftwerk Bremen in Germany. © Weserkraftwerk Bremen GmbH.

The newly built Three Gorges dam in China is often cited as a negative example of
the type of environmental damage hydropower plants can cause. Twenty cities and
more than 10 000 villages, together housing more than one million people, were
sacrificed for the building of this plant. It is still not clear what the ecological impact
has been on the flooded area. It is expected that the many environmental sins of the
past will come back to haunt those who live in the area, for example by contaminat-
ing the groundwater. The fear is that the 600 km-long reservoir is turning into a
cesspit of sewage and industrial waste.
On the other hand, when it is finished, the Three Gorges dam will produce 84 billion
kilowatt hours of electricity per year. This equates to more than one-sixth of total
German demand. If modern coal-fired plants were used to generate the same amount
of electricity, they would be emitting more than 70 million tons of carbon dioxide
into the atmosphere each year. This corresponds to the total carbon dioxide emis-
sions of Austria.
In this sense it is important to weigh the benefits and disadvantages of all hydro-
power plants. It is possible to build ecologically viable plants. The main thing is
that ecological aspects are included in the equation along with protecting the envi-
ronment and providing reasonably priced energy.

9.7 Hydropower Markets

About 17% of the electricity generated worldwide comes from hydropower plants.
In 2004 Canada was still the leader in this field (Figure 9.17), but in the meantime
it has been overtaken by China. The hydroelectric share of the power generated
varies substantially from country to country. Whereas 100% of Norway’s electricity
comes from hydropower plants, this share is 84% in Brazil, 65% in Austria and

207
9 Hydropower Plants – Wet Energy

61% in Canada. The hydroelectric share for China and the USA is only 16% and
6%, respectively. At 4% Germany’s share is relatively insignificant. In Europe,
Norway, Iceland and Sweden have the highest percentage of hydroelectric energy,
followed by the Alpine countries.

others
1055.0 billion
kWh

Japan 77.4
India 99.0 China
397.0

Norway
134.4

Russia
172.9 Canada Figure 9.17 Generation of electricity
359.9 from hydropower plants in different
USA
263.8 Brazil countries. Status 2005. Data: (US
334.1 Department of Energy DOE, 2008).

The large-scale power plants have a massive output. The Itaipu facility in Brazil,
with the highest output in the world (Figure 9.18), generates more electricity than
all hydropower plants in Germany and Austria combined.

Figure 9.18 Aerial view of the Itaipu power plant. Photo: Itaipu Binacional, www.itaipu.gov.br.

208
9.8 Outlook and Development Potential

9.8 Outlook and Development Potential

Among the different types of renewable energy generation, the use of hydroelectric
power is the most developed. In the industrialized nations most of the potential has
already been exploited. The potential for new large-scale plants now mainly exists
in the developing and emerging countries. Classic run-of-river and storage hydro-
power plants could at best double the output of electricity generated worldwide.
However, in the long term their share of energy production worldwide will decrease,
because the demand for electricity will continue to rise.
Certain types of hydropower plants, such as wave and ocean current plants, which
are not yet in use, show some promising potential for the future. However, substan-
tial cost reductions are still needed before these types of power plants can make any
major impact on the market.
The biggest advantage of hydroelectric energy is its relatively consistent output of
power compared to solar energy and wind power. This aspect makes it easier to
plan a mix of different renewable energies. With the share of renewable energies
being used to generate electricity increasing, storage and pumped-storage power
plants are also becoming important options due to their ability to offer consistency
in electricity supply.

209
10
10 Geothermal Energy – Power
from the Deep

When planet earth came into being four billion years ago, its form was consider-
ably different from what it is now. At that time earth was in a partially melted
state. It was not until about three billion years ago that the temperature of the
earth’s surface dropped to below 100 °C and the earth’s crust gradually began to
harden.
It may not seem like it in the depths of a northern winter, but today our planet is
anything but a cold ball. About 99% of the earth is hotter than 1000 °C and 90% of
the rest has temperatures of over 100 °C. Fortunately for us, these high temperatures
are almost exclusively found in the earth’s interior. Every so often volcanoes
produce impressive eruptions spewing molten matter from depths of up to 100 km
(Figure 10.1). Different technologies of plutonic geothermal energy enable us to tap
the heat of the earth’s interior in a controlled way so that we can satisfy some of
our heat and electricity demands.

10.1 Tapping into the Earth’s Heat

Earth itself is made up of concentric (Figure 10.2) bands comprising the core, the
mantle and the crust. The earth’s core has a diameter of around 6900 km (4290
miles). A differentiation is made between the outer, liquid core and the inner core,
which is made of solid matter. Maximum temperatures in the earth’s core reach
6500 °C, which is hotter than the surface of the sun.

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

210
10.1 Tapping into the Earth’s Heat

Figure 10.1 Volcanic eruptions bring the energy from the earth’s interior to the surface in a spectacular
fashion. Photos: D.W. Peterson and R.T. Holcomb, US Geological Survey.

Figure 10.2 The structure of the earth.

211
10 Geothermal Energy – Power from the Deep

Why the Earth’s Inner Core is Not Liquid


The earth’s inner core, which essentially consists of iron and nickel, has
temperatures of up to 6500 °C. Under normal environmental conditions with
an ambient pressure of one bar, these temperatures would make iron and nickel gaseous.
The pressure rises as the depth into the earth’s interior increases. This pressure reaches
maximum values of four million bars. This extremely high pressure ensures that the
outer core is liquid and the inner core is solid.

Eighty percent of the earth’s core consists of iron. A small amount of the heat in
the earth’s interior comes from residual heat from the time earth was formed but a
major part consists of radioactive decaying processes. The earth’s mantle is around
2900 km thick. We are only able to reach the top part of the earth’s crust.
The earth’s crust and the uppermost part of the earth’s mantle form the lithosphere.
The thickness of the lithosphere varies from a few kilometres to more than 100 km.
It consists of seven large and some smaller lithosphere plates (Figure 10.3). These
rather brittle plates float in the asthenosphere, where matter is no longer solid. The
plates are constantly in motion. Earthquakes and volcanoes frequently occur in areas
where two plates collide. Thermal anomalies can also frequently be observed in
these areas. High temperatures can occur there even at shallow depths, creating
conditions that enable us to make particularly effective use of the earth’s heat.

Figure 10.3 Tectonic plates on earth. Source: US Geological Survey.

212
10.1 Tapping into the Earth’s Heat

Regions like Central Europe that are not near the borders of tectonic plates do not
have optimal geothermal resources. This does not mean that high temperatures do
not exist underground. However, compared to regions like Iceland where geother-
mal conditions are favourable, it takes drilling at greater depths to reach the same
temperatures.
The best geothermal conditions in Germany are in the lowland plains of the Rhine
(Figure 10.4). In this region temperatures of 150 °C or more can be found at depths
of 3000 m. The average thermal depth gradient is around 3 °C per 100 m. On this
basis, a temperature increase of 90 °C could be expected at depths of 3000 m. In
Iceland temperatures like this occur at depths of just a few hundred metres.

Figure 10.4 Temperatures in Germany at depths of 1000 m and 3000 m. Graphics: www.liag-hannover.de
(Schellschmidt et al., 2002).

Deep drilling is required to exploit the high temperatures. The technique for this
type of drilling has long been familiar from oil extraction. Rotary drilling methods
are employed, whereby motors drive a diamond bit into the depth. At very great
depths, a drill bit cannot be driven in the conventional way over a drill string due
to the increased strain caused by turning and friction. An electric motor or turbine
therefore directly drives the drill bit.
The actual derrick (Figure 10.5) that holds the drill pipe is all that can be seen
from the surface. Water is pressed into the borehole at pressures of up to 300
bars through the inside of the driller. This flushing of the mud forces crushed rock

213
10 Geothermal Energy – Power from the Deep

material in the outer area between the bit and the borehole to the surface and at the
same time cools the drill. The drilling drive can also be used to change the bore
from its vertical position. This enables the boreholes that are sited close together on
the surface to exploit a larger area underground.

Figure 10.5 Left: New and used drill bits. Right: Structure of a derrick. Photos: Geopower Basel AG.

Depending on the conditions underground, the walls of a borehole can become


unstable. Steel tubes are inserted at large intervals and stabilized with a special
cement to prevent the walls from collapsing. The drilling is then continued with a
smaller drill bit. For many years the very high salt content of thermal water caused
serious problems at many drilling sites. Salt attacks metal and very quickly causes
corrosion. Today the use of specially coated materials has eliminated this problem.
The deepest drilling ever carried out was for research purposes on the Kola penin-
sula in Russia. This site had a depth of 12 km. In Germany the continental deep
geothermal drilling in the Oberpfalz region reached a depth of 9.1 km. These drilling
depths represent the technical limit today. The conditions at depths of 10 km are
extreme, with temperatures of more than 300 °C and pressures of 3000 bars. Under
these conditions rock becomes plastic and viscous.
Much shallower depths are required for geothermal use. The maximum drilling
depth currently planned for large-scale sites is around 5 km. However, the tech-
niques employed even at these depths are complicated and therefore expensive.
A distinction is made between the following in the exploitation of geothermal
deposits:

214
10.2 Geothermal Heat and Power Plants

■ Hot steam deposits


■ Thermal water deposits
■ HDR (hot dry rock)
Hot steam and thermal water sources can be used directly for heating purposes or
to generate electricity. If the underground consists only of hot rock, this can heat
up cold water that is compressed into the depth.

10.2 Geothermal Heat and Power Plants

10.2.1 Geothermal Heat Plants


If boreholes already exist in a thermal water area, it is comparatively easy to develop
a geothermal heat supply. In geothermal heat plants a feed pump fetches hot thermal
water from a production well to the surface (Figure 10.6). As thermal water often
has a high salt content, along with certain natural radioactive impurities, it cannot
be used directly to provide heat. A heat exchanger extracts the heat from the thermal
water and transfers it to a district heating grid. A reinjection well injects the cooled
thermal water back into the earth.

Figure 10.6 Principle of


a geothermal heat plant.

Relatively low temperatures of 100 °C or less are sufficient for heating purposes.
Deep drilling depths are therefore not necessary.

215
10 Geothermal Energy – Power from the Deep

A central heat system controls the amount of output depending on the heat require-
ment. If heat requirements are particularly high, a peak-load boiler can cover the
heat peaks. A backup boiler is also helpful to guarantee reliable heat supply in case
problems occur with the extraction pump or the well.

10.2.2 Geothermal Power Plants


Using geothermal energy to generate electricity is somewhat more complex than
providing thermal heat. For one thing, the relatively low temperatures that power
plant technology requires for thermal energy present new concepts such as:
■ Direct steam use
■ Flash power plants
■ ORC (Organic Rankine Cycle) power plants
■ Kalina power plants
Normal steam turbine systems can be used in geothermally optimal locations with
temperatures between 200 °C and 300 °C. If hot steam deposits exist underground,
they can be used directly to drive the turbines.
If hot thermal water is under pressure, it can be evaporated through an expansion
stage. The steam from the water that is still hot can in turn be transferred directly
to a steam turbine. This technique is called a flash process.
At temperatures of 100 °C or less geothermal water is not hot enough to vaporize
water. A normal steam turbine using water as the work medium is not suitable in
this case. An ORC (Organic Rankine Cycle) system is used instead (Figure 10.7).

Figure 10.7 Principle of a geothermal ORC system.

A steam turbine also forms the core of this kind of system. However, instead of
water, the steam turbine uses an organic material such as Isopentan or PF5050. A
heat exchanger transfers the heat from the geothermal cycle to the organic working
fluid. This material also evaporates under high pressure at temperatures lower than

216
10.2 Geothermal Heat and Power Plants

100 °C. The steam from the organic material drives the turbine and expands in the
process. The organic material is liquefied again in a condenser. A cooling tower
dissipates the residual heat. A feed pump again puts the organic material under
pressure and the heat exchanger completes the cycle.
The disadvantage of ORC systems is that they are relatively inefficient. At tempera-
tures of 100 °C the efficiency is well below 10%. This means that at best 10% of
the geothermal heat energy can be converted into electric energy. The Kalina
process promises slightly higher efficiency. With this process a mixture of water
and ammonia acts as the working material. However, not even the laws of physics
can make this process a viable option. Generating electricity from geothermal
energy is basically preferable and offers higher efficiency.
The first geothermal power plant in Germany is in Neustadt-Glewe between
Hamburg and Berlin (Figure 10.8). This ORC plant, which was built in 2003,
has 230 kilowatts of electric power. A geothermal heat plant, which was put
into operation in 1994 with heat output of 10.4 megawatts, is located at the
same site. The heat plant mainly generates electricity during the warm part of
the year when there is a surplus of geothermal heat due to the low demand for
thermal heat.

Figure 10.8 The Neustadt-Glewe geothermal heat power plant was the first plant to generate
electricity from geothermal energy in Germany.

In late 2007 the second geothermal power plant in Germany was put into operation
in Landau. It uses two boreholes to access 155 °C hot thermal water from a depth
of 3000 metres. An ORC process designed for continuous all-year-round power
generation uses this supply to generate around 2.5 megawatts of electricity.
The next power plant was built a short time later in Unterhaching near Munich.
Hot thermal water with a temperature of 122 °C is extracted here from a depth
of 3350 m. This geothermal cogeneration plant with an electricity output of

217
10 Geothermal Energy – Power from the Deep

around 3.6 megawatts is the first one to use the Kalina process. In addition,
the plant feeds heat with a thermal capacity of 28 megawatts into a district
heating grid.
Other countries such as the USA have been using geothermal power for more than
50 years and have much larger capacities. The Geysers, 116 km north of San
Francisco, is the largest dry steam field in the world. The construction on The
Geysers began in 1960. Today 15 power plants there have a net electricity-
generating capacity of about 725 megawatts, enough to power a city the size of San
Francisco.

■ iga.igg.cnr.it International Geothermal Association


■ geothermal.marin.org Geothermal Education Office

10.2.3 Geothermal HDR Power Plants


At drill depths of up to 5000 m temperatures are around 200 °C, even in regions that
do not benefit from optimal geothermal conditions, such as Germany, France and
Switzerland. As a result, fully acceptable efficiency is achieved in power generation
provided the depth is great enough.
Thermal water cannot usually be exploited at such great depths. What is mainly
found is HDR (hot dry rocks). Artificial shafts are created in which water can be
heated to extract the heat from the rocks. These shafts are made by compressing
water at high pressure into a borehole. The heat expands the borehole, creates new
fissures and expands existing cracks. This produces an underground fracture system
that can extend to several cubic kilometres. A listening borehole monitors the
activities.
This direct tapping of hot water deposits is also referred to as hydrothermal
geothermal energy, whereas hot-dry-rock technology is also called petrothermal
geothermal energy.
To generate geothermal energy, a pump transports cold water through an injection
well into the depth. There the water disperses into the cracks and fissures of the
crystalline rock and heats up to temperatures of 200 °C. The hot water reaches the
surface again through production wells and delivers the heat over a heat exchanger
to a power plant process and district heating grid (Figure 10.9).
During the 1970s the Los Alamos laboratory in the USA conducted the first tests
of the HDR method. A European research project on HDR technology has been
underway in Soultz-sous-Fôrets in Alsace since 1987. Its goal is to implement a
pilot power plant. In 2004 Geopower Basel AG was founded in Switzerland with
the aim of building the first commercial HDR power plant, but work on the project
was stopped in 2007 when small tremors occurred during the creation of the under-
ground fractures.

218
10.3 Planning and Design

Figure 10.9 Diagram


of an HDR power
plant.

10.3 Planning and Design

The most important factor in planning geothermal power plants is the temperatures
that can be achieved. The design of the heat exchangers, district heating grids and
plant processes are all based on projected temperatures. Geologists try to determine
in advance at which depths the desired temperatures can be found. To an extent they
are able to rely on the knowledge gained from existing boreholes.
In addition to achievable temperatures, extractable water quantities also play a major
role. Large volumes of water are needed to achieve high output. The diameter of
the borehole as well as the pumps both have to be designed accordingly. Last but
not least, it is important that the temperature of the thermal water is not allowed to
drop too significantly during the extraction process. Large-scale power plants
usually extract more heat from the depth than regularly flows back into an exploited
area. Therefore, a slow cooling of the exploited area cannot really be prevented.
The goal is to plan the intervals between the drilling so that the desired temperatures
can be sustained for about 30 years. After this period the temperatures drop below
the desired target values, and, as a result, the geothermal plant performance also
decreases. A new site, which should not be more than a few kilometres from the
existing site, must be developed if any further exploitation is planned.

219
10 Geothermal Energy – Power from the Deep

10.4 Economics

Drilling is by far the biggest cost factor with deep geothermal energy. Yet it is not
only the cost of the drilling itself that is the problem. The risks associated with it,
especially on commercial projects, are something that should not be underestimated.
Even the best geologists can never accurately predict what conditions will be like
underground. Drilling costs shoot up immediately if drills unexpectedly hit crystal-
line hard rock instead of soft sedimentary rock. If, in addition, the temperatures
underground are well below what was projected, a geothermal project can fail at an
early stage.
Underground conditions can often spring other surprises. For example, during work
on a geothermal project in Speyer in Germany, drillers not only found the thermal
water they were hoping for but also discovered an oilfield 2000 m below the surface.
So in addition to accessing hot thermal water of 160 °C, further drilling will also
exploit this oilfield.
Even when drillings are successful, geothermal power plants have to spend up to
half of all investment on the drilling activities themselves. Therefore, the cost of
geothermal power is often considerably higher than that of electricity from wind
and hydropower systems.
The renewable energy law in Germany also promotes geothermal power generation.
In 2007 the feed-in-tariff for geothermal power plants with a power capacity of up
to 5 megawatts was 15 cents per kilowatt hour. As this compensation was not high
enough to motivate the building of large numbers of commercial plants, a planned
revision to the renewable energy law is to raise this compensation to 20.5 cents per
kilowatt hour. This compensation includes a bonus for simultaneous power-heat
coupling and for petrothermal systems; that is, HDR (hot dry rock) power plants.
With new plants this compensation will drop by 1% per year.
The costs in countries with good geothermal conditions are considerably lower than
those in Germany. At drilling depths of a few hundred metres the drilling costs in
those areas end up being very low. If, in addition, high temperatures exist close to
the surface, some countries such as Iceland can even use geothermal heat to keep
their pavements free of ice in winter (Figure 10.10).

10.5 Ecology

Geothermal cogeneration plants have little ecological impact. Most of the plant is
located underground and is not visible, and therefore does not have a direct negative
impact on people or landscape. Only the power plant complex is above ground. Like
other heat plants, these also require cooling water for the plant processes. However,
water is readily available at most geothermal sites.
What can cause problems are some of the working materials used in the generation
of power. For instance, the material PF5050 used in ORC processes has a very high

220
10.6 Geothermal Markets

Figure 10.10 The Nesjavellir geothermal power plant in Iceland. Photo: Gretar Ívarsson.

greenhouse potential. If 1 kg of this material reaches the atmosphere, it produces


the same greenhouse effect there as 7.5 tons of carbon dioxide. However, alterna-
tives, such as Isopentan, are available.
In the long term geothermal systems can cause a certain cooling of some localized
areas under the ground. According to the knowledge available today, this has no
effect on the surface.
Relatively little research has been done on the risk of seismic activities. Small
tremors with an intensity of up to 3.4 on the Richter scale occurred in 2006 after
geothermal drilling for an HDR power plant in Basel, Switzerland. During the drill-
ing, water had been compressed at depths close to 5000 m. The tremors caused small
cracks to buildings in the region, so work on the project was halted. The geothermal
company involved has paid compensation for most of the damage.
As long as scientists are unable to predict accurately whether and when tremors can
occur during the compression of water, HDR projects will pose a certain risk in
densely inhabited regions. Hydrothermal geothermal projects that do not require
fissures and cracks to be artificially created are comparatively safe in terms of any
earthquake risk.

10.6 Geothermal Markets

China, the USA, Iceland and Turkey are the outright leaders when it comes to
geothermal heat use. The USA and the Philippines have the highest power plant
output for geothermal electricity generation (Figure 10.11). Providing more than
50% of the country’s energy needs, geothermal energy in Iceland constitutes the

221
10 Geothermal Energy – Power from the Deep

highest relative share of any country’s total primary energy supply. However, as
Iceland only has around 300 000 inhabitants and is not particularly densely popu-
lated, its absolute installed capacity is still lower than that in other countries.

10000
MW
9000
others
8000
Iceland
7000
New Zealand
6000 Japan
5000 Italy

4000 Indonesia
Mexico
3000
Philippines
2000
USA
1000
0
1990 1995 2000 2005

Figure 10.11 Installed geothermal power plant capacity worldwide. Data: IGS, iga.igg.cnr.it.

10.7 Outlook and Development Potential

Many countries are starting to develop the use of geothermal energy, but compared
to other renewable energy technologies, such as wind power and photovoltaics, the
annual growth rates for this energy source are quite modest.
The share of geothermal energy in the worldwide energy supply is therefore cur-
rently very low. But this technology has great potential. Another advantage of
geothermal energy is its constant availability. Compared to the fluctuations in certain
renewable energy sources – such as solar energy, wind power and hydropower –
geothermal energy is not subject to any unpredictable daily or yearly changes in
available supply. This makes geothermal energy an important building block in a
carbon-free energy supply. As the share of renewable energies increases in overall
energy demand, supply reliability will also become a big factor. This will help to
raise interest in building new geothermal plants.
For economic reasons, countries with large geothermal resources will remain the
leaders in the field. As prices for fossil fuels continue to rise, geothermal energy
will also become more interesting to countries with moderate geothermal resources.

222
11
11 Heat Pumps – from Cold to Hot
Due to the steep rise in oil and petrol prices during recent years and rising public
awareness of climate problems, alternatives such as wood pellet heating, solar
thermal systems and heat pumps are becoming more prominent. Manufacturers of
heat pump systems have seen a boom in sales since 2000.
Yet the whole principle behind the heat pump is far from modern. Lord Kelvin, a
British physics professor, already proved the principle of the heat pump in 1852.
He also recognized that a heat pump uses less primary energy to provide heat than
a system that heats directly. A heat pump uses a heat source with low temperatures
and increases it to a higher level of temperature (Figure 11.1). This process requires
an electric, mechanical or thermal drive.

Figure 11.1 Energy flow with a heat pump process.

11.1 Heat Sources for Low-Temperature Heat

A heat pump is basically a machine in which a mechanically or electrically driven


pump generates heat from a low-temperature source. This heat is then used to
provide heating or to produce hot water. Before a heat pump can even function, a

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

223
11 Heat Pumps – from Cold to Hot

low-temperature source must be available. The higher the temperature level of the
heat source, the more efficiently the heat pump can work.
The following heat sources are available to houses (Figure 11.2):
■ Groundwater (water/water)
■ Ground, ground heat exchanger/collector (brine/water)
■ Ground, earth probing device (brine/water)
■ Ambient air (air/water).
The waste heat from industrial plants can also be used.

Figure 11.2 Heat


sources for heat
pumps. Illustration:
Viessmann Werke.

Depending on the heat source, heat pumps fall into the categories of air/air, air/
water, brine/water or water/water systems. The heat medium supplied is indicated
in front of the oblique. In the case of ambient air, it is air. In the case of constantly
frost-free groundwater, it is water. Because of the risk of frost, a mixture of water
and antifreeze, called brine, flows through the pipes in the ground.
The transmitted heat medium is given after the oblique. In most cases, heat pumps
heat up water for heating and domestic use. They are seldom used to heat the air
for hot-air heating systems.
The higher the temperature of the heat source and the lower the temperature
needed for heating, the less electric energy is required to drive the heat pump. Due
to the low heating temperatures, underfloor heating is preferable to conventional
radiators.

224
11.1 Heat Sources for Low-Temperature Heat

Coefficient of Performance (COP) and Annual Coefficient


of Performance
The ratio of transient transmitted heat flow Q heating to transient, usually electric, drive
power P is called the coefficient of performance ε:
Q heating Q heating
ε= = 
P Qheating − Q source

The abbreviation for coefficient of performance is COP. The power output P and the
refrigerating capacity Q source of the low-temperature heat source together produce
the heat flow Q heating :
Q heating = P + Q source
If, for example, a heat pump with electrically driven power P = 3 kW produces heat
flow of Q heating = 9 kW , the coefficient of performance is ε = 3. The difference of
Q source = 6 kW derives from the low-temperature heat source.
The coefficient of performance only applies to transient values. What is interesting
is the annual average. This is called an annual coefficient of performance.

A high annual performance coefficient is essential for the ecological and economical
operation of a heat pump. With an annual performance coefficient of 4, for example,
a heat pump can cover a heating requirement of 10 000 kilowatt hours per year using
2500 kilowatt hours of electric energy. With an annual performance coefficient of
2, the requirement for electric energy rises to 5000 kilowatt hours.
Very good systems reach annual performance coefficients of about 4. In practice,
the values are often below this. Table 11.1 shows typical annual performance coef-
ficients for different types of heat pumps from a field test in Southern Germany.

Table 11.1 Typical annual performance coefficients for electric heat pumps (Lokale Agenda-
Gruppe 21 Energie/Umwelt in Lahr, 2007).

Heat pump Heat source Annual performance Annual performance


coefficient with coefficient with
underfloor heating radiators
Brine/water Ground 3.6 3.2
Water/water Groundwater 3.4 3.0
Air/water Air 3.0 2.3

The heat pumps that draw their heat from the ground produced the best values. The
annual performance coefficients for groundwater heat pumps were somewhat lower.
The reason is that it takes more pumping effort to extract groundwater than it does
to exploit heat from an enclosed brine loop in the ground. Furthermore, dirt traps

225
11 Heat Pumps – from Cold to Hot

eventually accumulate in wells for groundwater, which further increases the amount
of pumping energy needed. As ambient air temperatures in winter are lower than
ground or groundwater temperatures, hot-air heat pumps work most efficiently at
that time of the year.

11.2 Working Principle of Heat Pumps

All heat pumps need a refrigerant that is located in a closed loop. The refrigerant
absorbs the low-temperature heat. The heat pump then heats up the refrigerant to a
higher temperature, the heat of which is then used. Based on the working principles,
a distinction is made between
■ Compression heat pumps
■ Absorption heat pumps
■ Adsorption heat pumps

11.2.1 Compression Heat Pumps


Compression heat pumps are by far the most common type (Figure 11.3). The
principle of these heat pumps is based on a refrigerant with a low boiling point that
vaporizes when temperatures are very low and reaches high temperatures when it
is compressed (Table 11.2). The heat supplied from the low-temperature source in
the vaporizer is sufficient for vaporization.

Figure 11.3 Principle of a compression heat pump.

A compressor (usually electrically driven) brings the vapour-forming refrigerant to


a high operating pressure. During this process it heats up considerably. This process
is similar to what happens with a bicycle pump when one uses one’s thumb to stop

226
11.2 Working Principle of Heat Pumps

Table 11.2 Temperature ranges for popular refrigerants.

Abbrev. Name Boiling point Condensation


at 1 bar temperature at 26 bars
R12 Dichlordifluormethane −30 °C 86 °C
R134a 1,1,1,2-Tetrafluorethane −26 °C 80 °C
R290 Propane −42 °C 70 °C
R404A Mixture of different HFC −47 °C 55 °C
R407C Mixture of different HFC −45 °C 58 °C
R410A Mixture of different HFC −51 °C 43 °C
R600a Butane −12 °C 114 °C
R717 Ammonia −33 °C 60 °C
R744 Carbon dioxide −57 °C −11 °C
R1270 Propene −48 °C 61 °C

the air escaping while energetically pumping a tyre. The heat of the heated refriger-
ant is then used as useable heat, usually for room heating or heating water. The heat
is removed through a condenser that again liquefies the refrigerant. The refrigerant
that is compressed expands over an expansion valve, cools off and is transferred to
the vaporizer again.

The Turned-Around Refrigerator


Heat pumps are also used in refrigerators. They act like cooling machines.
A vaporizer removes the heat from the interior of a refrigerator. The heat is
emitted over cooling fins on the back of the appliance. The heat emitted there comprises
the heat removed from the refrigerator and the electric driving energy of the refrigerator
compressor. This explains why a room with an open refrigerator door cannot be cooled
down in the summer. A refrigerator emits more heat from the back than it removes from
the inside.
The first compression refrigerating machine was developed by the American Jacob
Perkins in 1834. He used ether, which is no longer used today, as a refrigerant for his
ice-making machine. The refrigerant ether has the disadvantage that in combination with
atmospheric oxygen it forms a highly explosive mixture. This occasionally caused ether
ice machines to explode.

11.2.2 Absorption Heat Pumps and Adsorption Heat Pumps


Like compression heat pumps, absorption heat pumps use low-temperature heat
to vaporize a refrigerant. However, absorption heat pumps use a thermal com-
pressor instead of the electrically driven compressor of compression heat pumps
(Figure 11.4).

227
11 Heat Pumps – from Cold to Hot

Figure 11.4 Principle of absorption heat pumps.

The function of a thermal compressor is to compress and heat the refrigerant. This
happens through a chemical process of sorption, for example through the dissolving
of ammonia in water. This was explained in the section ‘Cooling with the sun’ in
Chapter 6. The heat released through sorption can be used as thermal heat.
A solvent pump transports the solution to the generator. The solvent pump, unlike the
compression heat pump, does not build up high pressure, so a relatively low amount of
electrically driven energy is needed. The generator now has to separate the water and
refrigerant ammonia in the solution again to enable the sorption to take place once more.
High temperature heat is needed for the expulsion. Solar heat and biogas can be used.
The high temperature heat supplied is well below the dissipated quantity of useful
heat. The main advantage of absorption heat pumps is that they use a very small
amount of valuable electric energy. They are particularly useful for large-scale
applications. They are also used as refrigerating machines in refrigerators run with
propane gas. The refrigerant ammonia is toxic and flammable. However, it is a
widely used chemical and is considered easy to control.
Adsorption heat pumps, which only differ by two letters from absorption heat
pumps, likewise use thermal energy as the driving energy.
What is meant by adsorption is that a gas like water vapour is added to a solid
material, such as active coal, silica gel or zeolite. The process of adsorption, thus
the bonding of the water vapour through the solid material, creates high tempera-
tures that can be utilized by a heat pump. However, adsorption heat pumps are still
at the research stage and therefore will not be covered in detail in this book.

11.3 Planning and Design

Today heat pumps are available for almost any heat capacity desired. Manufacturers
usually provide advice on selection and design. The most important aspect in plan-
ning is the selection of a low temperature heat source and how to exploit it.

228
11.3 Planning and Design

If a heat pump is to be installed in a water conservation area, no groundwater is


allowed to be drawn. Approval for the use of a deep-ground probe to utilize the heat
from the ground is only given in special cases as the brine could contaminate the
groundwater if there is a leak. In Switzerland, for example, there are ground probes
in water conservation areas where carbon dioxide (R744) replaces the brine.
Using ambient air as a heat source is the simplest and most cost-effective approach.
This is not even a problem in water protection areas. No approvals are necessary to
install and operate air/water or air/air heat pumps in these areas. All these heat
pumps basically need are two openings in a house wall through which the ambient
air can be fed to the heat pump (Figure 11.5). Condensation forms if the outside
air is very cold and should drain off in a controlled way. Heat pumps can also
easily be installed outdoors. Air/water heat pumps function at ambient temperatures
as low as −20 °C. A supplementary electric heater helps to ensure that necessary
heating requirements are covered when temperatures are particularly extreme. A
small buffer storage unit can optimize the operating times of a heat pump. The
disadvantage of using ambient air is that the annual performance coefficients are
relatively poor. Compared to other heat sources, it necessitates a higher use of
electricity.

Figure 11.5 A heat-pump system. Source: BBT Thermotechnik GmbH.

Brine/water heat pumps, in other words heat pumps that extract heat from the
ground, are the types that use the least amount of energy. Either ground collectors
or ground probes are used to extract the heat. A ground collector usually comprises
a set of plastic pipes that are laid in a snakelike pattern in the soil (see Figure 11.2).
The optimal depth for the pipes is 1.2 to 1.5 m, and the gap between pipes should
be about 80 cm.

229
11 Heat Pumps – from Cold to Hot

Size of Geothermal Heat Collector

The length l and the area A of a ground collector is calculated from the required
refrigerating capacity Q source of the low-temperature heat source and the extraction
capacity q per metre of pipe and the pipe gap dA:
Q source
l= and A = l ⋅ d A .
q
If the heat capacity desired, for example, Q heating = 10 kW , and the COP ε = 4, a
refrigerating capacity of Q source = 7.5 kW is required. With dry sandy soil the extrac-
tion capacity is around q = 0.01 kW m ; with dry clay soil, 0.02 kW/m. As a result,
on the basis of this example, the length of pipe needed for clay soil is calculated as
7.5 kW
l= = 375 m
0.02 kW m
And with a pipe gap of dA = 0.8 m the collector area is
A = 375 m ⋅ 0.8 m = 300 m 2 .
If any doubt exists, it is recommended that the values be rounded off generously.
As individual pipe lines should not exceed lengths of 100 m, the recommendation
is that four circles of pipe, each 100 m in length, be used.

If sufficient space is not available in a garden or there is no desire to dig up a whole


plot of land, ground probes can be used to reach the ground heat. Vertical boreholes
can reach depths of 100 m. Constant temperatures of around +10 °C or more exist
at these depths all year round. U-shaped pipe probes are inserted into the boreholes
through which the brine of the heat pump will later flow. The depth of the drilling
and the number of probes depend on the heat requirement and the composition of
the ground below. Geologists and specialized drilling firms can help with the speci-
fications. Depending on the composition of the ground at the bottom, the potential
extraction capacity is between 20 and 100 watts per metre. A rough estimate could
be calculated at about 55 watts per metre. Therefore, about 5.5 kilowatts of refrig-
erating capacity could be extracted from a 100 m-deep probe. Several parallel probes
with a gap of at least 5 to 6 m between them are necessary to achieve higher capaci-
ties (Figure 11.6).
In addition to the ground heat, heat from groundwater can be used. This requires a
production well and a reinjection well. The reinjection well transfers the cooled
groundwater back into the ground. It should be sited at least 10 to 15 m behind the
production well, in the direction of the groundwater flow, so that the cooled water
does not flow back to the supply drilling.
In many countries approval must be obtained from the responsible water authority
before groundwater can be extracted. This approval is usually given under certain

230
11.3 Planning and Design

Figure 11.6 Air/water source heat pump installed outdoors and not requiring drilling (left). Deep drilling
for a water/water source heat pump (right). Photos: STIEBEL ELTRON.

conditions, except in water conservation areas. Permission is also often required to


drill the ground probes for closed brine/water systems. Normally, the drilling
company applies for the required approvals.

From the Idea of a Heat Pump to Owning One’s Own System

■ Determine possible heat sources:


Ground probe – Does deep drilling require approval?
Ground collector – Can a large section of the garden be dug up?
Groundwater – Is the house located in a water conservation area?
Air – Final solution if other sources are not available.
■ Calculate heat requirement and heating capacity.
Can insulation help to reduce the heat requirement?
Can the required temperatures be reduced through underfloor heating or large
radiators?
■ Request quotations for heat pumps and, if necessary, for the drilling.
Tip: Only HFC-free heat pumps are optimal for protecting the climate.
■ If necessary, have the drilling company obtain approval for drilling.
■ Determine optimal energy tariff, and, if necessary, plan to have buffer
storage unit.
Tip: Only green energy is optimal for protecting the climate.
■ Arrange for system to be installed by a qualified company.

231
11 Heat Pumps – from Cold to Hot

11.4 Economics

The heat pump sector has seen high growth rates in recent years, mainly for eco-
nomic reasons. In 2009 the investment costs in Europe for a typical heat pump for
a single-family home were between 8 000 and 12 000 euros. The costs of developing
a heat source, which are in the order of 3000 to 6000 euros for ground collectors
or ground probes, are additional to this.
The costs of conventional heating systems are not applicable to new buildings with
heat pumps. For example, with gas heating, the costs can include a gas connection
and a chimney in addition to the gas burner. Nevertheless, the investment costs for
heat pump systems are usually considerably higher than for conventional gas or oil
heating.
The advantage of heat pumps is the low running costs. Compared to conventional
gas heating, there is no expenditure for gas meters, maintenance and chimney clean-
ing. The fuel prices, and hence also the expenditure on energy to operate a heat
pump, are usually lower than in the case of oil and natural gas. The fuel price
advantage with heat pumps depends partially on the annual coefficient of perform-
ance and energy tariffs. With electricity bought at a regular tariff and a moderate
annual coefficient of performance of 3, the fuel costs for a heat pump could be even
higher than those for natural gas and oil. With a good annual performance coefficient
of 4, a heat pump offers advantages in fuel costs even based on the standard energy
tariffs available since 2005 (Figure 11.7).

0.08

0.07

0.06
fuel price in €/kWh

0.05

0.04

0.03

natural gas
0.02
f uel oil
electricity, standard tarif f COP=3
0.01
electricity, standard tarif f COP=4
electricity, special tarif f COP=4
0.00
1996 1997 1998 1999 2000 2001 2002 2003 2004 2005 2006 2007 2008

Figure 11.7 Development of domestic prices for natural gas, crude oil and electricity for
the operation of heat pumps for different coefficients of performance (COP) in Germany.

A number of energy companies offer special tariffs for heat pumps. However, the
operation of a heat pump often has to be interrupted during the times of the day
when peak tariffs apply. A buffer storage unit is needed to bridge the heating

232
11.5 Ecology

requirements during these periods. The heat pump should also have additional
capacity to produce surpluses for these times. The special tariff is often one-third
less than the standard tariff. If a special tariff can be applied to the heat pump, the
fuel costs for the pump will normally be considerably lower than for natural gas or
oil heating. Users who want to use climate-compatible green energy to run a heat
pump will usually not benefit from any special tariffs.
Whether a heat pump pays for itself in the long term depends mainly on the devel-
opment of gas, oil and energy prices. Biomass heating shows evidence of low fuel
prices and is another alternative to oil and natural gas heating.

■ www.heatpumpcentre.org IEA Heat Pump Centre


■ www.ehpa.org European Heat Pump Association

11.5 Ecology

Heat pumps are generally regarded as a positive use of the environment. But this is
not always the case. The Achilles heel of heat pumps is the refrigerant. The range
of refrigerants for compression heat pumps is broad. Chlorofluorocarbons (CFC)
were often used during the first heat pump boom. Because of their negative impact
on the ozone layer, their use has been banned in new systems.
Today hydrofluorocarbons (HFC), often called CFC equivalent substitutes, are
mostly used. Although they are harmless to the ozone layer, they share another
characteristic with CFC that is negative for the environment: both materials have
extremely high greenhouse potential. As a result, even small quantities of between
1 and 3 kg of refrigerant in heat pumps for single-family homes are developing into
an ecological problem.
If 2 kg of HFC R404A reaches the atmosphere, it develops the same effect on the
climate there as 6.5 tons of carbon dioxide. This quantity of carbon dioxide is
emitted during the burning of 32 500 kilowatt hours of natural gas. The same amount
heats a standard new-build house for three years, and a three-litre house for no fewer
than nine years. The energy needs of the heat pump itself are not even included in
this balance sheet.
If a leakage occurs in a heat pump system, the refrigerants escape quickly because
they evaporate under normal environmental conditions. On the positive side, HFC
substances are neither toxic nor inflammable. However, in terms of climate-
compatibility, the fast volatility of refrigerants turns out to be a problem. Not every
heat pump will develop a problem whereby its entire content of refrigerant escapes
into the atmosphere. However, refrigerant loss is unavoidable during the filling
and disposal of a system and, due to continuous seepage, during regular operation.
Nowadays standard heat pump suppliers seldom use refrigerants that can adversely
affect the climate. At the same time heat pumps with R290 or propane as the

233
11 Heat Pumps – from Cold to Hot

refrigerant are not showing performance data that is any worse that those with
refrigerants containing HFC. Special safety measures must be taken due to the flam-
mability of refrigerants R290, R600a and R1270, but for the most part these are
easy to implement. In the past more pressure to use HFC-free refrigerants was obvi-
ously placed on the manufacturers of refrigerators and freezers than on those of heat
pumps. As a result, in Europe refrigerants that are not harmful to the environment
have become part of the standard range in this area for years, despite their flam-
mability. In contrast, the HFC problem with heat pumps is hardly ever discussed.
Most manufacturers publicize the HFC they use as being environmentally friendly.
One manufacturer is even brazenly claiming on his homepage that the refrigerant
R407C is HFC-free. The layperson will find it almost impossible to distinguish
between appliances and devices that contain HFC and those that are HFC-free.
Table 11.3 provides some further guidance in this area.

Table 11.3 Greenhouse potential of different refrigerants relative to carbon dioxide.

Abbrev. Name Material group Greenhouse potential


R12 Dichlordifluormethane CFC 6640
R134a 1,1,1,2-Tetrafluorethane HFC 100
R404A Mix of different HFC HFC 3260
R407C Mix of different HFC HFC 1530
R410A Mix of different HFC HFC 1730
R290 Propane HFC-free 3
R600a Butane HFC-free 3
R744 Carbon dioxide HFC-free 1
R717 Ammonia HFC-free 0
R1270 Propene HFC-free 3

Suppliers of heating systems often list heat pumps in the ‘renewable energy’ cate-
gory. But this is only correct up to a point. Although most of the useable energy of
a heat pump is in the form of renewable low-temperature heat from the environment,
the power almost always comes from an electrical socket. This power is delivered
by regular energy suppliers that frequently offer special tariff conditions because of
the high quantity of electricity purchased for heat pump systems. In many countries
this electricity ends up coming from coal-fired or nuclear power plants. In Norway,
however, hydropower plants generate almost all the electricity used in the country.
This makes the heat pump there a completely renewable system. Some countries
offer the option of changing to green energy suppliers. In this case, too, a heat pump
system is completely renewable and therefore free of direct carbon dioxide emis-
sions. However, it usually means that suppliers of heat pumps have to dig a bit more
deeply into their pockets.

234
11.6 Heat Pump Markets

If conventional rather than green energy is used to operate a heat pump, the savings
in carbon dioxide emissions are still relatively low due to the poor efficiency of
fossil thermal power plants compared to modern natural gas heating (Figure 11.8).
If the heat pump also uses a HFC refrigerant that impacts the environment, in an
extreme case the environmental balance sheet can turn out to be even worse than
with a modern heating system using natural gas.

Figure 11.8 Environmental balance sheet on two heat pump heating options and natural gas heating.

11.6 Heat Pump Markets

After the first oil crisis in the 1970s, the heat pump sector experienced a real boom.
However, due to technical problems, a drop in oil prices and a lack of environmental
compatibility, the market for heat pumps collapsed almost completely by the late
1980s. The market did not revive until the mid-1990s and is currently experiencing
high annual growth rates. Figure 11.9 shows the sales of heat pumps for Germany
as an example of market development.
However, heat pumps are much more popular in certain other countries than in
Germany, and around 450 000 were installed in the European Union in 2006. With
122 000 systems installed, Sweden has the biggest heat pump market by a large
margin. In 2006, around 75% of all newly built single-family homes in Switzerland
had heat pump heating. As the average carbon dioxide emissions resulting from the
generation of electricity in Switzerland and Sweden are substantially lower than in
Germany, heat pumps in those countries also have a much more favourable envi-
ronmental balance sheet, in addition to cost advantages.

235
11 Heat Pumps – from Cold to Hot

70000

60000
sales of heating heat pumps
50000

40000

30000

20000

10000

0
1978
1979
1980
1981
1982
1983
1984
1985
1986
1987
1988
1989
1990
1991
1992
1993
1994
1995
1996
1997
1998
1999
2000
2001
2002
2003
2004
2005
2006
2007
2008
Figure 11.9 Sales of heat pumps for heating in Germany.

11.7 Outlook and Development Potential

The environmental balance sheet of heat pumps is improving continuously as a


result of the steady increase in the share of renewable energies used to generate
electricity. If in the medium term HFC-free alternatives start replacing refrigerants
that contain HFC, which cause climate problems, from the ecological point of view
the heat pump will develop into a very interesting alternative to conventional heating
systems.
As renewable heating systems, such as solar thermal systems and biomass heating
(see Chapter 12), can only cover a certain proportion of the heat requirement in
many countries, the heat pump is an important component for carbon-free heat
supply in the long term. Buffer storage can also be used to change the operating
times of heat pumps. This would enable heat pumps to be partially centrally control-
led, thus helping to reduce service peaks in the electricity network. With a high
availability of wind power, for example, heat pumps would fill heat storage tanks
and then draw the heat again at times when power supplies are low. These possibili-
ties indicate that a further expansion of the heat pump market can be expected.

236
12
12 Biomass – Energy from Nature
Man has been using energy from firewood for the last 790 000 years – ever since
Stone Age people discovered how to make fire (Figure 12.1). This makes biomass
the oldest renewable energy source by a huge margin. In fact, biomass was the most
important energy supply worldwide well into the eighteenth century. Even today
some countries like Mozambique and Ethiopia use traditional biomass to cover over
90% of their primary energy needs.
As the use of fossil energy supplies grew, biomass use was almost non-existent in
the industrialized nations. In 2000 the share of biomass in the primary energy supply
in countries such as Britain, Germany and the USA was not even 3% .

Figure 12.1 People have been using the energy from firewood for thousands of years.

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

237
12 Biomass – Energy from Nature

Biomass started to become popular again even in the industrialized countries when
oil prices began rising dramatically at the beginning of the twenty-first century. In
addition to its traditional use in the form of firewood, modern forms of biomass are
now being exploited. Biomass is not only used to kindle simple open fires but also
to operate modern heating systems and power plants for generating electricity as
well as to produce combustible gases and fuels.

12.1 Origins and Use of Biomass

The term ‘biomass’ refers to a mass of organic material. It comprises all forms of
life, dead organisms and organic metabolism products. Plants are able to create
biomass in the form of carbohydrates through photosynthesis. The energy needed
is supplied by the sun (Figure 12.2). Only plants carry out this process; animals can
produce biomass only from other biomass. This is why all animals would starve to
death without plants.

Figure 12.2 The sun is responsible for the growth of biomass on earth.

Origin of Biomass

Through photosynthesis plants convert carbon dioxide (CO2), water (H2O) and
auxiliary substances like minerals into biomass (CkHmOn) and oxygen (O2):
H2 O+ CO2 + auxiliary materials + energy → C
kH n + H2 O+ O2 +
O
m
biomass
metabolic products
In the simplest case, so-called oxygenic photosynthesis produces glucose (C6H12O6):
12 H2 O+ 6 CO2 + solar energy → C 6 H12 O6 + 6 H2 O+ 6 O2.

238
12.1 Origins and Use of Biomass

Almost all the oxygen in the earth’s atmosphere is formed through oxygenic pho-
tosynthesis. Therefore, the oxygen we need in order to breathe is a pure by-product
of biomass production.

Biomass is distributed in many different forms throughout earth. In addition to solar


energy, water is essential for the growth of biomass. Even the solar energy in the
northernmost regions of the world is sufficient to create biomass. Regions with water
shortages, however, have low biomass growth (Figure 12.3).

Figure 12.3 The view from space – vast areas of water cover the globe. Source: NASA.

Plants therefore convert sunlight into biomass using natural chemical processes. An
efficiency can also be established for this process. As a result, land usage for
biomass cultivation is comparable to other renewable energy technologies such as
solar systems. A plant’s efficiency is determined by dividing the calorific value of
dried biomass by the solar energy that reached the plant during its growth phase.
On average, including deserts and oceans, the efficiency of biomass production on
earth is 0.14% [Kleemann and Meliß, 1993]. Despite the comparatively low effi-
ciency, biomass is created worldwide with an energy content that corresponds to
about ten times our entire primary energy requirements.
Yet not all biomass can be used as energy. Human beings currently use around 4%
of new biomass. Two percent goes into food and fodder production and one percent
ends up as wood products, paper or fibre. Around 1% of newly created biomass is
used as energy – usually in the form of firewood – and it therefore covers about
one-tenth of the world’s primary energy demand.
The plants that reach the highest efficiency during the conversion of sunlight into
biomass are C4-plants. These plants have a rapid photosynthesis and, as a result,
use solar energy particularly effectively. C4-plants include amaranth, millet, corn,

239
12 Biomass – Energy from Nature

Table 12.1 Biomass potential in Germany (Kaltschmitt et al., 2003).

Useable quantity Energy potential


in millions of tons in PJ/a
Stalk-type (straw, grass) 10–11 140–150
Wood and wood residue 38–40 590–622
Biogas substrate (biomass waste and residue) 20–22 148–180
Purification and landfill gas 2 22–24
Energy plant mix 22 298
Total biomass potential 92–97 1198–1274

sugarcane and willow. Under optimal conditions these plants achieve efficiencies
of 2 to 5%.
With biomass a distinction is made between the use of waste from agriculture and
forestry and the selective cultivation of energy plants. In Germany studies have
showed there is a total potential of around 1200 petajoules per year (Table 12.1).
This equated to about 8% of Germany’s primary energy needs in 2005. The potential
should be similar for other industrialized countries with a high population density.
Even if extensive energy-saving measures are implemented, biomass can still only
cover part of the energy demand.
There are many different possibilities for biomass use (Figure 12.4). The biggest
potential exists with wood and wood products. Even waste from agriculture and
forestry and biogeneous waste are important for the energy sector. In addition to
waste use, special energy plants can be grown. However, as energy plants compete
with food production for arable land, there is some controversy over large-scale
cultivation of these plants.

Figure 12.4 Possibilities for biomass use.

240
12.2 Biomass Heating

The next step is processing the biomass materials. These materials are dried,
compressed, fermented into alcohol, converted into biogas, pelletized or processed
into fuel in chemical plants. The aim of the processing is to produce useable
biomass fuels.
This biomass fuel has the same spectrum of use as fossil fuels like coal, crude oil
and natural gas. Biomass power plants can use biofuels to generate electricity;
biomass heating can satisfy heating needs; and biofuels can be used to run cars and
other vehicles.
The versatility of biomass use has led to a real interest in alternative fuels. In many
industrialized countries like Germany and Britain, however, biomass falls far short
of becoming a complete replacement for fossil fuels. Nevertheless, biomass fuels
will play an important role in the renewable energy sector in the future.

12.2 Biomass Heating

With traditional biomass, the focus has been on the generation of heat for cooking
and heating. Even today the use of biomass for heating is one of its key applications.
Wood, straw and biogas are the commonly used fuels. Vegetable oils and bioalcohol
are also used in some heating systems.

12.2.1 Wood as a Fuel


Wood is by far the main fuel used for biomass heating. It is available in different
processed forms (Figure 12.5). As the first step, felled trees are cut to a common
length to produce round wood. High-quality woods are not used as fuel but are
processed further by the timber industry.
The round wood is then cut up, either by hand or by machine, to produce firewood.
Wood scraps or inferior-quality wood may be processed into wood shavings, which
can be made into wood briquettes or wood pellets. Special compression techniques
are used to press the wood into the right shape for burning. The natural lignin of
the wood serves as a binder, so no additional binders are required.
Because of their uniform shape and small size, wood pellets are an ideal fuel. They
can easily be delivered in bulk tankers and then blown into special pellet stores.
This eliminates the need for time-consuming manual loading. Automated feeding
systems enable wood pellet heating systems to provide the same level of heat and
ease of operation as natural gas or oil heating systems.
Some isolated quality problems existed in the early days of wood pellet production.
Pellets that are the wrong size can get stuck in conveyor systems. If the pellets have
not been compressed sufficiently, they can disintegrate too quickly and block up a
system. Therefore, wood pellets should comply with European standards such as
the Austrian Ö standard M 7135 or the German DIN plus standard. Wood pellets
must conform to the following specifications:

241
12 Biomass – Energy from Nature

Figure 12.5 Different processed forms of wood. From top left to bottom right: round wood, firewood,
wood briquettes, wood pellets.

■ Diameter: 5–6 mm; length: 8–30 mm


■ Minimum calorific value: Hi: 18 MJ/kg or 5 kWh/kg
■ Bulk stacked density: 650 kg/m3; raw density: larger 1.12 kg/l
■ Water content: less than 10%; ash content: less than 0.5%
■ Limit values for sulphur, nitrogen, chlorine and abrasion
The standards applied in the USA are generally less strict than those in Europe. The
new standard CEN/TS 14961 applies to all future production of wood pellets. A
pile of pellets weighing one ton takes up 1.54 cubic metres of space and has a calo-
rific value of 5000 kilowatt hours. This equates to a calorific value of approximately
500 litres of heating oil. Therefore, two kilos of wood pellets can replace one litre
of heating oil.

A Tree – the World Record Holder


An 80-year-old beech tree reaches a height of 25 m (82 ft). The crown of the
tree has a diameter of 15 m (49 ft) and contains about 800 000 leaves. If the
leaves were spread out lying next to one other on the ground, they would cover an area
of about 1600 square metres (0.4 acres). This beech tree supplies the oxygen needs of

242
12.2 Biomass Heating

ten people and in the process absorbs large quantities of carbon dioxide. The 15 cubic
metres of wood of this beech have a dry mass of 12 tons, of which around six tons are
bound to pure carbon. This corresponds to the carbon content of 22 tons of carbon
dioxide.
Yet even this massive beech is dwarfed by some tree species. The majestic sequoia
reaches heights of up to 115 m (377 ft), diameters of 11 m (36 ft) with a circumference
of over 30 m (98 ft) and volumes of 1500 cubic metres. Not only do trees reach an
impressive size, but they can also live to a very great age. Some pine trees live for 5000
years. There is a special type of pine where the rootstock can even survive for more than
10 000 years. New shoots that sprout from them ‘only’ live to about 2000 years. As a
result, trees are presumed to be the oldest, highest and heaviest living things on earth.

If one were to compress wood shavings into a cube with sides each one metre long,
the mass and calorific value of this cube would be considerably higher than a cubic
metre of wood pellets. The reason is that a relatively large amount of air is trapped
within a pile of pellets.
In technical jargon, the measure of capacity for one cubic metre of solid wood mass
without any gaps is also called a solid cubic metre. When round wood or firewood
is stacked very neatly, air gaps occur. The measure of capacity in this case is called
stacked cubic metre. When firewood is thrown loosely on a pile, the air gaps
increase. In terms of measure of capacity, this is referred to as bulk stacked cubic
metre (BCM). The various measurements of capacity can be converted approxi-
mately into one another, with the exact factors depending on the type of wood and
the shape the wood takes:
■ 1 solid cubic metre = 1.4 stacked cubic metre = 2.5 bulk stacked cubic metres
(BCM)
As we all know, wet wood burns poorly, because the calorific value of wood
depends critically on water content. Damp wood is also heavier than dry wood.
This means that not only the wood itself but also the water it contains has to be
transported. The water evaporates when the wood burns. However, to evaporate,
it needs energy – which in turn comes from the wood. The calorific value drops as
a consequence.
Either the wood moisture or the water content is used to indicate the drying level.
As both quantities have different parameters and their values therefore differ con-
siderably, this can lead to some confusion. The water content describes the propor-
tion of water in wet wood. On the other hand, the wood moisture indicates the mass
of water contained in ratio to the mass of totally dry wood. If exactly half the weight
of wood consists of water, the water content is 50% but the wood moisture level is
100%.
Completely dry wood with 0% water content is referred to as bone-dry wood. For
example, bone-dry beech trees have a calorific value of five kilowatt hours per kilo-
gram (kWh/kg). When wood is dried outdoors, its water content reduces to between
12 and 20%. The wood should be chopped as early as possible into small pieces,
covered and then left to air dry for at least one year, but ideally two years. Wood

243
12 Biomass – Energy from Nature

that has been allowed to dry in covered areas can even end up with water content
of less than 10%. Even with 15% water content, the calorific value of beechwood
is still 4.15 kWh/kg. With freshly cut wood with a water content of 50% the calorific
value drops to 2.16 kWh/kg (Figure 12.6). Consequently, its calorific value is con-
siderably less than half that of bone-dry wood.

6
beech, birch

5 pine, spruce
calorific value Hi in kWh/kg

1 wood moisture u in %
11 18 25 43 67 100 150
0
0 10 15 20 30 40 50 60
water content w in %

Figure 12.6 Calorific values of wood depending on wood moisture and water content.

This example shows that firewood should be well dried before it is burnt to extract
the optimal energy content. The mass-related calorific value per kilogram differs
minimally with different types of wood. On the other hand, the volume-related calo-
rific value, thus the calorific value of a solid cubic metre or a stacked cubic metre,
varies considerably (Table 12.2). Heavy wood like beech burns longer than light-
weight spruce.

Table 12.2 Characteristics of different types of firewood.

Calorific
value dried Density dried Calorific value Hi with w = 15%
Hi0 in In kg/solid In kWh/kg In kWh/solid In kWh/stacked
kWh/kg cubic metre cubic metre cubic metre
Beech 5.0 558 4.15 2720 1910
Birch 5.0 526 4.15 2570 1800
Pine 5.2 431 4.32 2190 1530
Spruce 5.2 379 4.32 1930 1350

In addition to providing poor calorific values, too much moisture in wood also has
some other undesirable effects. High water content means that wood is being burnt
under less than optimal conditions. As a result, it releases a high amount of harmful
substances and produces unpleasant quantities of smoke and a pungent smell.

244
12.2 Biomass Heating

12.2.2 Fireplaces and Closed Woodburning Stoves


The classic biomass heating system is the fireplace. For centuries open fireplaces
have been used to heat individual rooms. Yet this is a relatively inefficient use of
firewood. Open fireplaces usually only reach 20 to 30% efficiency. This means that
70 to 80% of the firewood’s energy escapes unused through a chimney. As romantic
as old castles and palaces may seem, the use of open fires to achieve consistently
pleasant room temperatures in draughty castle halls was an almost hopeless task.
With 70 to 85% efficiency, enclosed fireplaces and closed woodburning stoves are
considerably more effective than open fireplaces (Figure 12.7). A glass panel in
front of the stove can be opened to resupply it with wood and then closed after it
has been lit. In principle, fireplaces can also be used to heat up domestic water.
Fireplaces and closed woodburning stoves are normally only used as supplementary
heating systems because of the work involved in cleaning and lighting them, and
replacing the firewood.
One of the main problems with fireplaces is that they need a relatively large amount
of air. In addition to the air needed to enable a fire to burn well, large amounts of
unused air escape through the chimney. A fresh air supply from outdoors to burn a
fire is essential for well-insulated and airtight houses.

Figure 12.7 Enclosed fireplace and closed woodburning stove. Source: BBT Thermotechnik GmbH.

12.2.3 Firewood Boilers


The firewood boiler (Figure 12.8) is an option for those who want to heat with
reasonably priced firewood and without the bother of a fireplace or stove. As these
boilers are usually installed in a cellar, they do not have aesthetic merits of a
fireplace. They have large containers for wood supplies that need to be stocked
manually, and they can burn for several hours on a single load of wood.

245
12 Biomass – Energy from Nature

Figure 12.8 Solid fuel


boiler for heating with
firewood. Source: BBT
Thermotechnik GmbH.

Compared to the top burn-up of fireplaces where the flame rises upwards, many
boilers work with the burn-up at the bottom or on the side. Air is fed into the boiler
so that the flame is shown pointed either downwards or to the side. This increases
burning time and cuts down on emissions. The controls, which mainly regulate air
supply, ensure that a boiler is burning at an optimal level and adapt it to the heating
requirement. Firewood boilers are available in different performance classes and
reach maximum efficiencies of more than 90%. The efficiency of smaller boilers is
usually somewhat less.
As firewood boilers cannot be regulated downward to any specific temperature
desired, the installation of buffer storage is recommended. This will enable a boiler
to work under optimal operating conditions at all times. The storage absorbs the
excess heat and then continues to supply the heating needed after the firewood has
burnt down. The combination of a firewood boiler with a solar thermal system is a
good idea because the boiler can be switched off completely in summer when there
is very little need for heating.

12.2.4 Wood Pellet Heating


Wood pellet heating systems offer by far the greatest ease of operation. The fuel is
kept in a special pellet store, and an automated feed mechanism using either a feed-
ing screw or a suction device transports the pellets directly to the burner. A screw
conveys the pellets from the bottom part of the store. With a suction device similar

246
12.2 Biomass Heating

to a big vacuum cleaner the pellets are also sucked up from below. The suction
hoses are very flexible and even enable the bridging of large distances between the
store and the burner. As the suctioning of the pellets can be noisy, modern pellet
boilers have a small hopper from which the pellets are conveyed to the burner
through gravity or a small feeding screw. The hopper is then filled from the store
via an automatic timer switch so that the pellet feeding system does not disturb
anyone at night.
Optimally, the store would be in the cellar. It should be big enough to cover fuel
requirements for one year. If a cellar is not available, the pellets can also be
stored in special silos in a large utility room or in an adjoining shed. Waterproof
tanks in the ground are also suitable for stocking pellets. Normally, the store will
have two openings (Figure 12.9). The pellets are blown in directly through
one opening from the tanker truck that delivers them. The displaced air from
the blowing process escapes through the other opening. As the blowing-in of the
pellets creates a great deal of dust, a filter traps the wood dust before the air is let
out again.

Figure 12.9 Wood pellet heating with pellet store. Source: BBT Thermotechnik GmbH.

The controls of pellet boilers always ensure that an adequate supply of pellets is
available, thereby guaranteeing that the operation of the system is fully automated.
The flame is also lit automatically by an electric hot-air blower. When the required
heating level has been reached, the heating system switches off again independently.
Another feeding screw transports the accumulated ashes to a special ash pan. Wood
pellet heating therefore does not require any manual intervention for its daily opera-
tion. As with firewood boilers, buffer storage should be installed to reduce the
frequency with which a fire has to be lit.

247
12 Biomass – Energy from Nature

Wood pellet heating requires a very small investment of the user’s time. The soot
and ash residue has to be cleaned out of the pellet boiler room every second month
or so. In addition, the ash pan should ideally be emptied once or twice a year. The
ash can be used as fertilizer in the garden.
In addition to boilers designed for cellars, attractive pellet boilers are available for
use in living areas (Figure 12.10). These boilers can be installed in sitting rooms
and show a nice flame. Again, a suction device conveys the pellets from the store
that would still be located in the cellar.

Figure 12.10 Pellet boiler for living area use with pellet store in a cellar (left). Source: Windhager
Zentralheizung.

12.3 Biomass Heat and Power Plants

In addition to being burnt in stoves and boilers for heating systems in single homes
and apartment blocks, biomass can also be used in large heat plants. A central heat
plant consists of a high-performance boiler and a fuel store. The fuel store is usually
big enough to ensure that independent operation can be guaranteed for several days
or even weeks. A district heating grid then transports the heat to the consumers
connected to the grid (Figure 12.11).
With centralized heat plants, individual consumers no longer have to worry about
fuel procurement and system maintenance. These tasks are handled by the operator
of the heat plant. The efficiency of large heat plants is often somewhat higher than
that of small non-central systems. On the other hand, heat losses are higher because
of the long pipes in the district heating grid. However, large heat plants fare con-
siderably better when it comes to the emission of harmful substances. Compared to
heating systems in single-family homes, large plants use more modern filtering

248
12.3 Biomass Heat and Power Plants

Figure 12.11 The Altenmarkt biomass heating plant in Austria and district heat distributors.
Source: Salzburg AG.

techniques and have stricter conditions on emissions. This ensures that combustion
gases emit fewer harmful substances.
Electricity generation is another important application of biomass. Centralized
systems that are used exclusively to generate electricity are called power plants.
Biomass power plants function in a similar way to coal-fired power plants. The fuels
used include wood residue, wood shavings and straw. A steam boiler burns the
biomass and produces steam that drives a steam turbine and an electric generator.
The principle of steam power plants is explained in the section on parabolic trough
power plants in Chapter 7.
Compared to photovoltaic systems and wind turbines, electricity generated by
biomass plants is not dependant on prevailing weather conditions. Biomass fuels
are optimal for storage and can be used as and when needed. This makes biomass
power plants a viable supplement to other renewable energy plants. They can guar-
antee electricity supply in situations when there is little wind or sun available at the
same time.
In contrast to coal-fired plants that can output more than 1000 megawatts of power,
biomass plants have a considerably lower output in the order of 10 to 20 megawatts.
The use of biomass fuels is one of the main reasons for this. Biomass fuels usually
come from the region where the particular power plant is located. If the output of
the plant is too high, some of the biomass then has to be supplied from sources
much further away.
Numerous new biomass power plants have recently been built all over the world.
One example is the Königs Wusterhausen plant near Berlin, Germany (Figure
12.12). This power plant has an output of 20 megawatts and with 160 million kilo-
watt hours per year can cover the energy demands of around 50 000 households.
For fuel it uses 120 000 tons of waste and wood residues from the Berlin region
annually. The efficiency of this biomass plant is around 35%.

249
12 Biomass – Energy from Nature

Figure 12.12 The Königs Wusterhausen biomass power plant in Germany (left) and a wheel loader
for biomass fuel transport (right). Source: MVV-press photo.

Apart from its use in pure power plants, biomass is also appropriate for heat power
plants. In addition to electricity, these plants produce heat which district heating
grids distribute to consumers. Heat power plants generate both power and heat,
which is known as power-heat coupling. Heat power plants normally use biomass
fuel more effectively than pure power plants that only generate electricity. The
important thing is that a buyer can always be found for the heat that is generated.
Heat is not often needed during the summer months. During this time of the year a
heat power plant may operate less efficiently than a power plant that is purely opti-
mized to generate electricity.
Heat power plants are also built in lower output classes so that they can be suitable
for industrial buildings, houses and apartment blocks. As these plants are usually
offered in a modular configuration for a building-block style, they are referred to
as cogeneration units. However, the efficiency of smaller systems is frequently
lower than that of large centralized systems. In addition to solid fuels like wood
shavings and pellets, cogeneration units use biofuels and biogas.

12.4 Biofuels

Fluid and gaseous biofuels are more versatile than wood. In addition to generating
heat and electricity, biofuels can be used directly as fuel in the transport sector,
replacing petrol and diesel. Production methods are available that can convert dif-
ferent biomass raw products into biofuels. Unlike with food production, the prefix
‘bio’ in this case does not stand for controlled organic cultivation with minimum
effect on the environment. On the contrary, the raw materials for biofuels are usually
produced using conventional farming methods.

250
12.4 Biofuels

12.4.1 Bio-oil
The biofuel that is the easiest to produce is bio-oil. Over 1000 plants with a high
amount of vegetable oil are usable in the production of bio-oil. The most popular
ones are rapeseed oil, soya oil and palm oil (Figure 12.13). Oil mills produce the
vegetable oil directly through either pressing or extraction processes. The residue
from the pressing can be reused as animal feed.

Figure 12.13 Oil-rich plants such as rapeseed and sunflowers are raw material for vegetable oil.
Photo left: Günter Kortmann, North Rhine-Westphalia Chamber of Agriculture.

Very few older precombustion chamber diesel engines can be run on vegetable oil
without a problem unless they have been converted. Even engines like the Elsbett
engine, which were specifically developed to run on vegetable oil, have not yet
achieved any significant market share. Vegetable oil is somewhat tougher than diesel
fuel and needs higher temperatures to fire. However, even normal diesel engines
can be adapted and converted to run on vegetable oil.

12.4.2 Biodiesel
Biodiesel comes closer to having the characteristics of conventional diesel fuel
than pure vegetable oil. Vegetable oil and animal fat are the raw materials used
to produce it. The Belgian G. Chavanne applied for a patent for a method to
produce biodiesel as early as 1937. Chemically, biodiesel is fatty acid methyl
ester (FAME).
In Central Europe rape is normally used to produce biodiesel. Oil mills extract the
raw material rapeseed oil from rapeseed seedlings. The by-product rapeseed animal
meal usually ends up in the animal feed industry. Rapeseed oil methyl ester (RME)
is then created from the rapeseed oil in a transesterification facility.

251
12 Biomass – Energy from Nature

Production of Rapeseed Methyl Ester (RME)

For the production of RME, rapeseed and methanol together with a catalyst such as
a caustic soda solution are placed in a reaction vessel at temperatures of about 50 °C
to 60 °C. This produces the desired rapeseed oil methyl ester as well as glycerine:
⎯⎯→ glycerin + RME ( biodiesel )
rapseed oil + methanol ⎯catalyst

Biodiesel can be used as a substitute for fossil diesel fuels based on crude oil. It is
offered by many petrol stations in Germany. The engine manufacturer’s specifica-
tions should stipulate whether a vehicle can run on pure biodiesel. If an engine is
not designed for use with this fuel, there is a danger that the biodiesel will eventu-
ally destroy the hoses and seals and cause engine damage. Small quantities of
biodiesel can be mixed with conventional diesel without a problem, even without a
release from the manufacturer. The European biofuel directive of May 2003 stipu-
lates that countries across the EU must institute national measures to replace 5.7%
of all transport fossil fuels (petrol and diesel) with biofuels by 2010.
However, the positive environmental impact of biodiesel is not undisputed.

■ www.ebb-eu.org European Biodiesel Board


■ www.ieabioenergy.com IEA Bioenergy

12.4.3 Bioethanol
Sugar, or glucose, starch and cellulose are used to produce bioethanol. The raw
materials include sugar beet, sugar cane and grains (Figure 12.14). Sugar can be
fermented directly into alcohol. On the other hand, starch and cellulose must first
be broken down.

Production of Ethanol from Glucose

Glucose can be converted directly into ethanol in a hermetically sealed environment


through fermentation with yeast:
C 6 H12 O6 ( glucose ) ⎯fermentation
⎯⎯⎯⎯ → 2 CH3 CH2 OH ( ethanol ) +
2 CO2 ( carbon dioxide )

Carbon dioxide is a by-product of this reaction. As plants absorb carbon dioxide


again during their growth, this reaction is actually not emitting any greenhouse
gases. The result of the fermentation is mash with an ethanol content of around 12%.

252
12.4 Biofuels

Figure 12.14 Corn and other types of grains are raw materials for bioethanol production. Photos: Günter
Kortmann, North Rhine-Westphalia Chamber of Agriculture.

Raw alcohol with a concentration of over 90% is extracted through a distilling


process. Dehydration over molecular sieves then finally produces ethanol with a
high degree of purity. The waste from the ethanol production can be processed into
animal feed. The amount of energy required to extract the alcohol is relatively high.
If this energy comes from fossil fuels, then the effects of bioethanol on the climate
are not favourable. In extreme cases, the impact can even be negative.
Bioethanol can easily be mixed with petrol. An E number indicates the ratio of the
mixture. E85 means that the fuel content is 85% bioethanol and 15% petrol. In some
countries small quantities of bioethanol are added to petrol. This is not a problem
if the ethanol portion is 5% or less. Normal petrol engines can even be run with an
ethanol portion of 10% without requiring any modification. Engines must be modi-
fied for the use of ethanol for anything above that level.
In Brazil flexible-fuel vehicles are very popular. These automobiles can be filled
with different mixtures including an ethanol portion of between 0 and 85%.
Production facilities that use rye, corn and sugar beet as raw materials in bioethanol
production have also been established in other countries.

12.4.4 BtL Fuels


In the case of pure vegetable oil, biodiesel and bioethanol, only the parts of a plant
that are rich in oil, sugar or starch can be used to extract fuel. The second generation
of biofuels is aimed at overcoming this drawback. The abbreviation BtL stands for
‘biomass-to-liquid’ and describes the synthetic production of biofuels. Various raw
materials, such as straw, bio waste, wood residue and special energy-rich plants,
can be used in their entirety in this process. The result is a major increase in the
potential and possible land area yield for the production of biofuels.

253
12 Biomass – Energy from Nature

The production of BtL fuels is relatively complex. The first stage is a gasification
of the biomass raw materials. Through the addition of oxygen and steam a synthesis
gas consisting of carbon monoxide (CO) and hydrogen (H2) is produced at high
temperatures. Different gas purification stages separate out carbon dioxide (CO2),
dust and other impurities such as sulphur and nitrogen compounds. A synthesis
process converts the synthesis gas into fluid hydrocarbons.
The best-known synthesis process is the Fischer-Tropsch process developed in
1925. Named after its developers Franz Fischer and Hans Tropsch, this process is
carried out at a pressure of around 30 bars and temperatures above 200 °C using
catalyzers. During the Second World War this process was widely used in oil-poor
Germany to extract much sought-after liquid fuels from coal. A different procedure
then produces methanol from the synthesis gas and processes it further into fuel.
During the final production processing stage the liquid hydrocarbons are separated
into different fuel products and refined (Figure 12.15).

Figure 12.15 Principle of the production of BtL fuels.

BtL fuels have not yet reached a stage where they are ready for mass production.
Various companies are currently setting up prototype facilities for producing syn-
thetic biofuels. For example, Volkswagen and Daimler have reserved the brand
names SunFuel and SunDiesel and are collaborating with the manufacturing firm
Choren Industries. The main advantage of BtL fuels is that they can replace con-
ventional fuels directly without the need for any engine modifications. However,
BtL fuels are comparatively expensive because of the complex production
procedures involved.

12.4.5 Biogas
In addition to its use in the production of liquid fuels, biomass can also be used in
biogas plants to produce biogas. In this process bacteria ferments biomass raw
materials in a moist, hermetically sealed environment. The centrepiece of a biogas
plant is the heated fermenter (Figure 12.16). A stirring device mixes the substrate
and ensures that homogeneous conditions exist. The biological decomposition

254
12.5 Planning and Design

process mainly converts the biomass into water, carbon dioxide and methane. The
biogas plant captures the gaseous components. The biogas extracted in this way
consists of 50 to 75% combustible methane and 25 to 45% carbon dioxide. Other
components include steam, oxygen, nitrogen, ammonia, hydrogen and sulphur
hydrogen.

Figure 12.16 Biogas plant in a cornfield and view of interior with stirrer. Source: Schmack Biogas AG.

At further stages the biogas is purified and desulphurized. It is then stored in a gas
storage tank. The biogas yield varies considerably depending on the type of biomass
substrate. Whereas with cow dung the gas yield is around 45 cubic metres per ton,
with corn silage the yield is at least 200 cubic metres per ton.
Biogas is used mostly in internal combustion engines. Petrol engines and modified
diesel engines are both appropriate. If the engine drives an electric generator, it
can produce electric power from biogas. The waste heat from the engine is also
useable.
After further processing, biogas can be fed directly into the natural gas grid. This
requires the removal of any trace gases, water and carbon dioxide. Green gas
suppliers are now marketing feed-in biogas in some parts of Germany. Changing
over to this type of gas supplier will actively promote the expansion of biogas
production.

12.5 Planning and Design

The possibilities for using biomass are extremely diverse, and would merit an entire
book all to themselves. Therefore, this section confines itself to the planning and
design of firewood boilers and pellet heating systems. Both systems are especially
relevant to single-family homes.

255
12 Biomass – Energy from Nature

12.5.1 Firewood Boilers


Certain facts need to be determined before a firewood boiler is installed. Wood-fired
systems may save on carbon dioxide, but they release harmful substances such as
dust and carbon monoxide. As a result, the use of solid fuels is regulated in certain
cities and towns. The local chimney cleaning company can be contacted to clarify
these conditions. A firewood boiler also requires sufficient storage area for firewood
and a suitable chimney for operation.
If no impediments to the installation exist, the next step is deciding on the dimen-
sions of the firewood boiler system. The performance of the boiler should at least
comply with the rated building heat load, in other words the maximum expected
heat power requirement. A boiler has to run constantly when outdoor temperatures
are extremely low. It is best to have a boiler large enough so that it only has to be
filled up once a day under average heating conditions.

Boiler Output and Buffer Storage Size for Firewood Boilers

The minimal boiler rated output Q B results from the rated burning period TB of one
stocking of fuel and the rated building heat load QH :
6.4
Q B = Q H ⋅ .
TB
With a building heat load of 10 kW, which corresponds approximately to that of a
standard new single-family house, the rated output of a boiler based on a rated
burning period of 2.5 h is:
6.4 h
Q B = 10 kW ⋅ = 25.6 kW
2.5 h
The necessary volume VS of buffer storage can be approximated from the boiler
rated output Q B and the rated burning time TB:
1
VS = Q B ⋅ TB ⋅13.5 .
kWh
As a result, the buffer storage can absorb the total heat volume of a boiler with a
fully stocked burning area. A boiler rated output of 29 kW and a rated burning time
of 2.5 h results in a storage volume of
1
VS = 29 kW ⋅ 2.5 h ⋅13.5 = 979 l.
kWh

12.5.2 Wood Pellet Heating


The general guidelines on firewood heating can also be applied to wood pellet
heating. As pellet heating systems normally transfer fuel automatically to the burner,

256
12.5 Planning and Design

there is no loss of comfort due to the frequent firing-up of the system during the
day. The boiler rated output can therefore be designed specifically for the heat
energy needs of a building. A buffer storage facility is still useful as it prevents a
boiler from firing up too frequently and ensures that the heating is working at its
rated load most of the time. The system then burns at an optimal level and produces
a minimal emission of noxious substances.
Unlike firewood, wood pellets are usually stored in a special area within a building.
It is not possible for the pellets to be stored outdoors as they will absorb moisture
and may be damaged. An area of three to five square metres should normally be
sufficient as the pellet store for a standard single-family new-build home. Older
buildings need more space, whereas three-litre houses require considerably less.
The transport system for the pellets is on the floor of the store (Figure 12.17). A
slanted shelf will ensure that the pellets slide down to the transport system even
when the filling level is low. However, slanted shelves create a void that reduces
the useable storage volume to about two-thirds of the space volume. As the price
of wood pellets varies throughout the year, the store should hold at least a year’s
worth of fuel. Pellets can then always be bought at times when prices are low.

Figure 12.17 Cross-section of a wood pellet store.

Size of Wood Pellet Store

If the annual heat requirement QH of a house and the boiler efficiency ηboiler are
known, this can be used to calculate the necessary store volume VS including the
void:
QH
VS = .
2 kg kWh
⋅ ηboiler ⋅ 650 3 ⋅ 5
3 m kg

257
12 Biomass – Energy from Nature

If the annual heat requirement is not known, an estimate can be made on the basis
of 200 kWh per square metre of living area for an average building, 100 kWh/m2
for a standard new-build and 30 kWh/m2 for a three-litre house based on Central
European climatic conditions. The heat needed to provide hot water is added to this
quantity.
On this basis, the heat required for a standard new-build with 130 m2 of living area
is 13 000 kWh. If 2000 kWh is added for hot water, the total heating requirement is
15 000 kWh. With an average boiler efficiency of 80% = 0.8 the store volume is
then calculated as
15 000 kWh
VS = = 8.7 m 3 .
2 kg kWh
⋅ 0.8 ⋅ 650 3 ⋅ 5
3 m kg
A store with a floor area of 2 m × 2 m and a height of 2.17 m would be sufficient.
This store can hold pellets with a pellet volume of 2/3 · 8.7 m3 = 5.8 m3. The mass
of the pellets amounts to 5.8 m3 · 650 kg/m3 = 3770 kg. If more space than this is
available, the store can be made larger. The cost of delivery quantities of 5 tons or
more is generally lower than for smaller quantities. A store volume of around
11.5 m3 is necessary to store this quantity. It is important that optimal storage condi-
tions exist so that the pellets do not suffer long-term damage due to high air
humidity.

■ www.pelletheat.org Pellet Fuel Institute

Steps to Installing Biomass Heating

■ Determine type of fuel.


Cut wood/firewood – using one’s own wood.
Wood pellets – for a fully automated mode of operation.
■ Estimate heat requirement and heat output.
If necessary, base needs on previous heating system.
■ Can the heat requirement be reduced through insulation?
■ Dimension of store – is an adequate store available?
Is sufficient space available for buffer storage?
Is the chimney suitable for connecting biomass heating?
■ Clarify regulations concerning system operation and residue removal,
for example, by a chimney sweep.
■ Request quotations for biomass heating

258
12.6 Economics

■ Check into favourable financing conditions within the framework of other


climate-protection measures, check and apply for available grants.
■ Arrange for system to be installed by a registered company with expertise in this
technology.

12.6 Economics

Trying to predict the long-term economic development of biomass fuels compared


to fossil fuels is a bit like reading tea leaves. This is because of the fluctuation in
wood pellet prices during recent years (Figure 12.18). Whereas in 2003 the prices
for heating oil and wood pellets of a comparable calorific value were practically the
same, in 2005 oil prices soared by 50%. This caused a boom in the demand for
wood pellets, which the industry had a hard time meeting. In late 2006 the prices
for wood pellets were even higher than comparable crude oil prices for a short time.
The prices normalized a few months later and oil prices again rose sharply.

500 100

_
450 90

oil in €ct/l, gas in €ct/10 kWh


400 80
wood pellets in €/tons

350 70
300 60
250 50
200 40
150 30
wood pellets
100 20
f uel oil
50 10
natural gas
0 0
01 03

07 03

01 04

07 04

12 04

07 05

12 05

07 06

12 06

07 07

12 07

06 08

12 08

06 09

Figure 12.18 Comparison of end user prices for fuel oil, natural gas and wood pellets
in Germany. Data: German Federal Statistical Office and German Energy Pellets Association.

The potential for producing wood pellets is far from enough to supply the total
current heating market. If more and more customers start using wood pellets as fuel,
the result will inevitably be a rise in prices. However, as fuel oil prices will also
continue to move upwards in the long term, the price advantage of wood pellets
could be maintained at a rising level.
Whether wood pellet heating makes economic sense depends primarily on the price
difference compared to fuel oil and natural gas. An estimate for installing this kind
of heating is around 15 000 euros, which is considerably more than it would be for
oil or natural gas heating. However, with natural gas there is the additional cost of

259
12 Biomass – Energy from Nature

a gas connection with new buildings. The cost advantage of wood pellets compared
to oil and gas is in the lower running costs of pellet heating. Depending on usage
and the price development of fuels, wood pellet heating can be amortized in just a
few years. The fuel prices for log heating are lower than with pellet heating. The
operating costs are therefore even lower.
The demand for biofuels is clearly increasing worldwide due to high oil prices and
the simultaneous increase in energy demand. This is also putting additional pressure
on food prices. Grain and corn prices have reached new records in recent years.
This development has also raised some ethical issues. Is it right that the quantity of
food products being processed into biofuels is increasing when more and more
people are unable to afford basic food as it is? An alternative is offered by second-
generation biofuels such as BtL fuels and biogas, which can be produced from the
non-edible parts of plants.
The Renewable Energy Law in Germany regulates the feed-in-tariff rates for
biomass power plants. These tariffs show the current rates needed to operate com-
mercial biomass power plants. The tariffs can differ considerably depending on the
biomass fuel used and power plant output. The maximum plant output eligible for
compensation is 20 megawatts. In 2009 the feed-in-tariffs ranged from 4.16 cents
per kilowatt hour for a 5-megawatt power plant using methane up to 11.67 cents
per kilowatt hour for a 149-kilowatt biomass power plant. In addition, bonuses of
up to 16 cents per kilowatt hour are given for power-heat coupling, the use of renew-
able waste and innovative plant technology. The amount of the tariff is based on
the year a plant starts operation and is then valid for 20 years.

12.7 Ecology

Biomass has also come under fire for ecological reasons. For example, a farmer in
Indonesia who sets fire to a hectare of rainforest in order to grow palm oil for
biodiesel to sell to Europe or North America is certainly not helping to protect the
climate. Priority should only be given to the sustainable production of biomass
raw materials if biomass really is to offer a long-term ecological alternative to
fossil fuels.

12.7.1 Solid Fuels


As explained earlier, biomass use is carbon-neutral. During its growth biomass
absorbs as much carbon dioxide as it releases again when it is burnt. However, the
prerequisite is that the use of biomass is sustainable. Therefore, the amount of
biomass that is used should be no higher than what grows back again.
In many countries, solid biomass fuels such as wood and straw usually come from
forestry or grain farming in nearby areas. Although the felling of trees, their trans-
port and finally the processing into fuel creates indirect carbon dioxide emissions,
these emissions are comparatively low. For cut wood from the direct vicinity they

260
12.7 Ecology

are almost zero. If one takes into account the indirect carbon dioxide emissions that
result from the production and the transport of wood pellets in an overall balance
sheet, the carbon dioxide emissions from wood pellet heating are still around 70%
lower than those of natural gas and more than 80% lower than those of oil heating
(Figure 12.19).

Figure 12.19 Environmental balance sheet for the use of biomass fuels.

The harmful substances that build up when biomass is burnt create far more of a
problem than indirect carbon dioxide emissions. Whereas large biomass power
plants have sophisticated filtering systems, single-family homes mostly use their
heating systems without any filtering mechanism. The blackened chimneys of their
fireplaces testify to this. Even if the carbon dioxide balance sheet turns out to be on
the plus side, biomass heating can release all sorts of other harmful substances into
the environment if it is not burnt properly. In Germany today the emission of
harmful fine dust from wood-burning plants is already in the same order as that
of motorized street traffic. Yet there are clear differences depending on the type of
heating system used. Due to their poor efficiency, open fireplaces generally cause
particularly high emissions. Therefore, the use of open fires is banned in many
places.
When dry firewood or standardized wood pellets are used in modern boilers, the
emission values with the same heat output can be 90% lower than with a fireplace.
Different companies are now working on producing filters to prevent the emission
of dust from small systems.

261
12 Biomass – Energy from Nature

12.7.2 Biofuels
The ecological impact of biomass fuels is even more controversial than the harmful
emissions for which they are responsible. Tractors and farm machinery emit carbon
dioxide just by being turned on. Added to this is the amount of energy required
to produce fertilizer and pesticides. Nitrogen fertilizers increase nitrous emissions,
which are harmful to the environment. Even the processing of biomass raw materials
to produce biofuels is energy-intensive. Huge quantities of carbon dioxide are
created if the energy needed comes from fossil energy sources.
The real difficulty in evaluating biomass fuels and their ecological impact is illus-
trated by the example of biomass cultivation in tropical rainforest regions – a topic
that has recently been a target of criticism. When tropical rainforests are cleared to
cultivate biomass, considerable quantities of carbon dioxide are released through
the usual slash and burn methods. The subsequent cultivation of raw materials for
the extraction of biofuels then has a negative environmental impact for many years.
In other words, from the standpoint of the environment, it would have been better
to burn crude oil from the start.
On the other hand, the results of the production of bioethanol on existing farmland
in Brazil are better than in Germany or the USA. Factories in Brazil mostly burn
the sugarless residue of sugar cane and extract the energy from this for ethanol
production. As a result, ethanol production there is largely carbon-neutral. Other
countries use substantial quantities of fossil fuel to do the same thing. This can
totally cancel out the climate benefits of bioethanol.
Another critical point with biomass fuels is the limited development potential. If all
the available land in the world were devoted to growing biomass for biofuels, it
would still probably not be enough to allow biofuels to replace total oil require-
ments. The second-generation fuels would not totally invalidate this argument but
would at least defuse it. If one removes one’s own energy needs for fuel production,
the net yield per hectare with BtL fuels is around three times higher than with
biodiesel (Figure 12.20). The land utilization of solar systems is clearly more

biogas
gross
net
BtL

bioethanol

biodiesel

vegetable
oil

0 1000 2000 3000 4000 5000 6000


litre diesel equivalent per hectare

Figure 12.20 Fuel yield per hectare for different biofuels. (1 hectare = 2.47 acres).

262
12.8 Biomass Markets

efficient than this. In Germany and Great Britain a photovoltaic system with 15%
efficiency can generate around 495 000 kilowatt hours of electric power per year on
one hectare of land or 200 000 kilowatt hours per acre. This corresponds to a con-
verted diesel equivalent of more than 50 000 litres per hectare.

12.8 Biomass Markets

Biomass use varies widely in different parts of the world. Biomass is by far the
most important energy source in the poorest countries of the world and in some
countries even constitutes more than 90% of primary energy needs (Figure 12.21).
The reasons for this are mainly economic. Most people in these countries simply
cannot afford crude oil, natural gas or electricity from coal-fired power plants. In
most industrialized countries like Britain, Germany and the USA the biomass share
is well below 10%. The only exceptions are countries like Finland and Sweden,
which are densely wooded and sparsely populated.

Ethiopia 93
Congo 93
92
Tanzania
Mozambique 88
Nepal 85
India 39
Finland 19
China 19
Sweden 16
Austria 9
Denmark 8
France 5
Spain 3
USA 3
Germany 2
Great Britain 1
Russia 0.6
Saudi Arabia 0

0 10 20 30 40 50 60 70 80 90 100
biomass share of the primary energy demand in %

Figure 12.21 Biomass share of primary energy demand in different countries.


Status: 2001, Data: (International Energy Agency, 2003).

Biomass is used very differently by industrialized and developing countries. In many


developing and emerging countries biomass is still used in the traditional way.
These countries see the main use of biomass as firewood for cooking. Industrialized
countries, on the other hand, focus on the modern forms of biomass use. These
include comfortable biomass heating systems, biomass power plants and biomass
fuels. The proportion of modern biomass is also increasing in industrialized coun-
tries. For example, Sweden has set itself the goal of becoming independent of crude
oil by 2020 mainly through an increase in biomass use. The potential is lower in
more densely populated countries like Germany and Britain. But even in these
countries a 10% biomass share in primary energy requirements is conceivable in
the long term.

263
12 Biomass – Energy from Nature

Germany can be used as an example showing the possible development. Between


1997 and 2007 the biomass share of end energy demand in Germany tripled and
now makes up more than 5% of the total energy demand (Figure 12.22).

180
electricity
160
secondary energy supply in TWh

heat

140 bioethanol
vegetable oil
120 biodiesel

100 biological waste

80

60

40

20

0
1997 1998 1999 2000 2001 2002 2003 2004 2005 2006 2007 2008

Figure 12.22 Development of biomass use in end energy supply in Germany.


Data: (German Federal Ministry for the Environment, Nature Conservation and Nuclear Safety, 2007).

12.9 Outlook and Development Potential

As the economies of the developing countries grow, the contribution of traditional


biomass sources to their energy supply will fall. On the other hand, high oil prices
and pressure to act on environmental protection measures will drive the use of
modern biomass. Biomass is capable of directly replacing fossil energy sources
without needing other technologies to do so. However, the potential of biofuels is
not sufficient to replace oil completely. Furthermore, the eco-balance of some
biofuels is hotly disputed. There are also ethical concerns about the growing
competition with food production. BtL fuels and biogas could be an alternative on
a smaller scale.
Biomass power plants will become increasingly important in meeting electricity
demand, not least because these plants are able to compensate for some of the major
fluctuations in output provided by wind turbine and photovoltaic systems. Biomass
will also play a bigger role in providing heat energy. It will therefore remain one
of the most interesting renewable energy sources for the future. Throughout the
industrialized countries biomass has the potential to achieve a double-digit share of
energy supply. Further developments in modern filtering systems are needed to
address the problem of harmful substances and fine dust in biomass burning.
However, special attention should be given to the sustainable use of biomass raw
materials. This is the only way in which it can make a noticeable contribution
towards protecting the climate.

264
13
13 The Hydrogen Industry and
Fuel Cells
Generations of schoolchildren have been entertained by the oxyhydrogen reactions
in chemistry class. When hydrogen is oxidized by oxygen from the air, the hydrogen
gas explosively releases its stored energy. A spark is enough to ignite a mixture of
hydrogen and normal air. In contrast to many other combustion processes, however,
the reaction product is absolutely harmless from an environmental point of view.
Hydrogen and oxygen simply react to pure water.
The proportion of electricity used in supplying energy is increasing all the time.
Fuel cells can generate the much sought-after electric power from hydrogen. The
only waste created is water. It is no wonder that the many people with a vision of
a global hydrogen economy see it as the solution to our current climate problems.
Hydrogen as a single energy source could at the same time help us to get rid of air
pollution, acid rain and other environmental problems caused by the use of energy.
Jules Verne saw the potential of hydrogen as early as 1874, and the question is why
a thriving hydrogen industry has not already developed. The answer is simple:
hydrogen does not occur in a pure form in nature. Energy and a complex technical
process are needed before it can be burnt again. This makes hydrogen expensive,
and some production processes involved even create high greenhouse emissions.
But even though a hydrogen industry is still only on the drawing board, and in the
long term would certainly never satisfy total energy demand, it is still an interesting
alternative energy source for some areas of application.

Jules Verne (1828–1905): ‘The Mysterious Island’

‘And what will they burn instead of coal?’ asked Pencroft. ‘Water,’ replied
Harding. ‘I believe that water will one day be employed as fuel, that hydrogen
and oxygen which constitute it, used singly or together, will furnish an inex-
haustible source of heat and light, of an intensity of which coal is not capable.
Water will be the coal of the future.’

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

265
13 The Hydrogen Industry and Fuel Cells

13.1 Hydrogen as an Energy Source

Hydrogen is by far the most common component in our solar system and constitutes
around 75% of the mass and more than 90% of all atoms. Our sun and the large gas
planets Jupiter, Saturn, Uranus and Neptune consist primarily of hydrogen. Here on
earth hydrogen occurs much less frequently. Its share of the total weight of earth is
only about 0.12%. Although hydrogen occurs more frequently in the earth’s crust,
it practically never occurs even there as a pure gas. Hydrogen is almost always
chemically bonded. The most frequent compound is water.
Hydrogen is the smallest and lightest atom. As an extremely light gas, hydrogen
was used to fill the gas bags of airships like the Zeppelins during the first half of
the nineteenth century. The Hindenburg disaster, where an electrostatic charge sup-
posedly caused the hydrogen to ignite, brought a tragic end to the prospects of
hydrogen use.
The main application of hydrogen today is in the chemical industry. As an energy
source it is currently used on a large scale mainly in the aviation sector and in
space travel. Hydrogen has occasionally been used to drive the jet engines of aero-
planes. In space travel liquid hydrogen is used as rocket fuel. For example, the
launch of a space shuttle consumes about 1.4 million litres of liquid hydrogen
weighing more than 100 tons. This is burnt along with the 0.5 million litres of liquid
hydrogen that the shuttle carries with it. The combustion temperature is up to
3200 °C (Figure 13.1).

Figure 13.1 Hydrogen is by far the most common element in our solar system (left) and is already
being used as an energy source for the launch of space shuttles (right). Graphic/Photo: NASA.

266
13.1 Hydrogen as an Energy Source

13.1.1 Production of Hydrogen


Hydrogen must first be produced in a pure form before the energy from it can be
used (Figure 13.2). This requires an easily available inexpensive raw material con-
taining hydrogen. Aside from water (H2O), which consists of hydrogen (H) and
oxygen (O), hydrocarbon compounds can also be an option. This is primarily natural
gas, or methane (CH4). Heating oil and coal consist of hydrogen (H) and carbon (C)
but have a much higher proportion of carbon than natural gas.

Figure 13.2 Procedure for producing hydrogen.

Current industrial methods for producing hydrogen almost exclusively use fossil
fuels, such as natural gas, crude oil or coal, as the raw material. Methods such as
steam reforming or partial oxidation to produce hydrogen from fossil hydrocarbons
chemically separate the carbon. It then reacts to carbon monoxide (CO), which can
be used energetically. The end product is carbon dioxide (CO2). These methods for
producing hydrogen are not real options for actively protecting the climate.
The Kværner method also uses hydrocarbons as the base material. However, the
waste product that it produces is active coal – therefore, a pure carbon. A direct
formation of carbon dioxide can be prevented with this method if the carbon is not
burnt further.
Basically, all the methods mentioned to produce hydrogen from fossil energy
sources are run at high processing temperatures. This requires large amounts of
energy. If this energy comes from fossil sources, this will again lead to the emission
of carbon dioxide. For climate protection it is usually better to burn natural gas or
oil directly than to take the circuitous route of producing hydrogen and then using
it as a supposedly environmentally friendly fuel.

267
13 The Hydrogen Industry and Fuel Cells

Other methods are therefore necessary to produce hydrogen so that it is environ-


mentally safe. The ideal way is with electrolysis. The German chemist Johann
Wilhelm Ritter first used electrolysis to produce hydrogen as early as 1800. Using
electric energy, electrolysis decomposes water directly into hydrogen and oxygen.
If the energy comes from a renewable energy power plant, the hydrogen can be
extracted free of carbon dioxide.
Alkaline electrolysis is an example (Figure 13.3). With this method two electrodes
are dipped into a conductive watery electrolyte. This can be a mixture of water and
sulphuric acid or potassium hydroxide (KOH). The anodes and cathodes conduct
direct current into the electrolytes. There they electrolyze water into hydrogen and
oxygen.

Figure 13.3 Principle of alkaline electrolysis.

Although electrolysis has already reached a high state of technical development as


an environmentally compatible option for oxygen production, other alternative
methods are also in development.
Thermo-chemical methods are an example. At temperatures above 1700 °C water
decomposes directly into hydrogen and oxygen. However, these temperatures
require expensive heat-resistant facilities. The required temperature can be reduced
to below 1000 °C through different coupled chemical reactions. For example, con-
centrating solar thermal power plants can produce these temperatures, and this has
already been successfully proven.
Other procedures include the photochemical and photobiological production of
hydrogen. With these procedures, special semiconductors, algae or bioreactors use
light to decompose water or hydrocarbons. These methods are also still at the
research stage. The main problem is developing long-term stable and reasonably
priced facilities.

268
13.1 Hydrogen as an Energy Source

13.1.2 Storage and Transport of Hydrogen


Once hydrogen has been produced, it has to be stored and transported to the con-
sumer. In principle, we are familiar with the storage and transport of combustible
gases from the use of natural gas. Hydrogen is an extremely lightweight gas with
very minimal density but has a relatively high calorific value. Compared to natural
gas, hydrogen with the same energy content requires much larger storage volumes
although the stored hydrogen is much lighter.
Hydrogen can either be compressed, stored under high pressure or liquefied in order
to reduce the necessary storage volumes. Under normal pressure hydrogen con-
denses but not until it reaches extremely low temperatures of minus 253 °C. A high
amount of energy is needed to achieve such low temperatures. Twenty to 40% of
the energy stored in the hydrogen is used to liquefy it. Liquid hydrogen is abbrevi-
ated as LH2.
In principle, the same technologies used in the natural gas sector can be used for
the liquidization, transport and storage of hydrogen. Hydrogen can be transported
either in pipelines or in special tankers and freighters. Whereas pipelines usually
transport the gaseous form, tankers are preferred for liquid hydrogen to reduce the
volume. In contrast to hydrogen, natural gas already becomes liquid at minus 162 °C
and is abbreviated as LNG (liquid natural gas) (Figure 13.4).

Figure 13.4 Experience from the natural gas sector can be used for storing and transporting hydrogen.
Left: Liquid gas tanker. Right: Pipeline. Source: BP, www.bp.com.

Compressed gas or liquid gas tanks are used to store small quantities. Another
disadvantage of hydrogen is the very small atomic portion, which makes it extremely
volatile. Large quantities of hydrogen can be lost if is stored in metal tanks for long
periods of time, because it diffuses through the storage walls. Underground storage
is recommended for large quantities of hydrogen, whereby it is compressed at high
pressure into underground storage shafts and extracted when needed.

269
13 The Hydrogen Industry and Fuel Cells

13.2 Fuel Cells: Bearers of Hope

Fuel cells are considered the key to the future energy use of hydrogen, because they
can convert hydrogen directly into electric energy. Theoretically at least, this results
in higher efficiency levels than with combustion in conventional thermal power
plants.
The principle of fuel cells has been known for a very long time. There is some
controversy about who actually invented the fuel cell. The German-Swiss chemist
Christian Friedrich Schönbein conducted the first tests in fuel cell technology in
1838. The English physicist Sir William Robert Grove built the first fuel cell in
1839. Well-known scientists like Henri Becquerel and Thomas Edison were subse-
quently involved in its further development. A sufficiently advanced stage of devel-
opment was finally reached in the mid-twentieth century, enabling NASA to make
major use of fuel cells by 1963.
Since the 1990s fuel cell development has been moving ahead at full speed. Car
manufacturers and heating companies have adopted the technology and are looking
to profit from a positive image as a result.
Fuel cells basically involve a reversal of electrolysis. A fuel cell always contains
two electrodes. Depending on the type of fuel cell, pure hydrogen (H2) or a fuel
containing hydrocarbons is fed through the anode and pure oxygen (O2) or air as an
oxidation material is fed through the cathode. An electrolyte separates the anode
and cathode (Figure 13.5). As a result of this, the chemical reaction is controlled.
Electrons flow over a large circuit and emit electric energy. The remaining positively
charged ions diffuse through the electrolyte. The waste product is water.

Figure 13.5 Working principle of fuel cells.

There are different types of fuel cells that essentially differ from each other based
on electrolytes, the permissible fuel gases and operating temperatures. In practice,
the following abbreviations are used to identify the fuel cell types:

270
13.2 Fuel Cells: Bearers of Hope

■ AFC alkaline fuel cell


■ PEFC polymer electrolyte fuel cell
■ PEMFC proton exchange membrane fuel cell
■ DMFC direct methanol fuel cell
■ PAFC phosphoric acid fuel cell
■ MCFC molten carbonate fuel cell
■ SOFC solid oxide fuel cell
Figure 13.6 shows the respective fuel gases and oxidation materials as well as the
electrolytes and operating temperature ranges for the different types of fuel cells.

Figure 13.6 Differences between fuel cell types.

The proton exchange membrane fuel cell (PEMFC) is the one most frequently used
today. In this fuel cell the electrolyte consists of a proton-conductive polymer film.
The fuel gases flow through carbon or metal substrates which serve as electrodes.
The substrates have a platinum coating that acts as the catalyzer. The typical operat-
ing temperature is about 80 °C. These cells do not require pure oxygen for operation
but can also work with normal air.
Because hydrogen as an energy source is only available in limited quantities today,
there is an interest in using fuel cells directly with energy sources like natural gas
and methanol that are relatively easily available. At a preliminary stage a reformer
uses a chemical process to break down hydrocarbons such as natural gas into hydro-
gen and other components. In the process a hydrogen-rich reformat gas is formed
from which a gas purification stage is still needed to eliminate harmful carbon
monoxide (CO) for the fuel cell.

271
13 The Hydrogen Industry and Fuel Cells

Molten carbonate fuel cells (MCFC) and solid oxide fuel cells (SOFC) work at much
higher temperatures. This enables gases containing hydrocarbons, such as natural
gas and biogas, to be used directly without any previous reforming. The dis-
advantage of high temperatures is the long time required for starting up and
switching off.
As the electric voltage of a single cell with values of around one volt is too low
for most applications, a number of cells are usually connected in series to form a
so-called stack (Figure 13.7).

Figure 13.7 Prototypes for fuel cells.

The electric efficiency of fuel cells today is usually in the order of 40 to 60%. Values
of more than 60% can be reached in individual cases. A fuel cell can therefore only
convert a part of the energy contained in hydrogen into electric energy. Basically,
the waste heat of fuel cells is also useable. Power-heat coupling, or the simultaneous
generation of power and heat, raises the overall efficiency of a fuel cell and can
increase it to over 80%.
Major advances have been made in fuel cell technology in recent years, and many
companies are already offering commercial units. However, the number of units
currently being sold is still relatively low. The price of fuel cell systems is still fairly
high compared to other energy supply units. Furthermore, it will be necessary to
increase the sometimes quite short service life of fuel cells if they are to have a
broader appeal.

13.3 Economics

It currently costs about 4 cents to produce a kilowatt hour of hydrogen through the
steam reforming of natural gas – assuming that natural gas prices are relatively
reasonable. One litre of petrol has a calorific value of about 10 kilowatt hours.

272
13.4 Ecology

Added to this would be the equivalent of hydrogen to one litre of conventional petrol
at about 40 cents. At this point the hydrogen would not yet even have reached the
tank for end user use. Including liquidization, transport and storage, the cost rises
threefold from the given price to well over one euro. This makes it more than double
the net petrol price in Europe.
Producing hydrogen in an environmentally compatible way through electrolysis
using renewable energy is even more expensive. With a prototype facility using
wind power for electrolysis, the equivalent of hydrogen to one litre of conventional
petrol would cost 5 euros. With large-scale technical plants, a price of 2 euros could
easily be achieved at the refuelling pump. Taxes and duties would have to be added
to this.
The long-term hope is that hydrogen at top renewable sites using electricity from
wind turbines, hydropower plants or solar thermal power plants will be available
for delivery at the equivalent of less than 2 euros per litre of petrol. If the petrol
price then rises well above 2 euros per litre, hydrogen at the pump would become
competitive. Presumably it will be some years before we get to that stage.
In addition to its use as a fuel, hydrogen is also considered an option for the large-
scale technical storage of electric energy. As relatively extensive losses always
occur during the electrolysis, storage and reverse conversion of hydrogen (Figure
13.8), this type of use based on the current state of technology is only viable in
isolated cases.

Figure 13.8 Losses when hydrogen is used to store electric energy, based on the current
state of technology.

This also applies to generating electricity from hydrogen that is produced in distant
areas and has to be transported to the consumer. The use of hydrogen in the electric-
ity sector will only be economically viable if major technological advances are made
in this area and power-heat coupling is used.

13.4 Ecology

The broad public perception of hydrogen as an energy source and fuel cells is
favourable, mainly because water is the waste product when hydrogen is used.

273
13 The Hydrogen Industry and Fuel Cells

However, what is important for the balance sheet on the environment is not what
comes out at the end but what is put in at the beginning. When steam reforming is
used to produce hydrogen from natural gas, around 300 g of carbon dioxide are
created per kilowatt hour of hydrogen (g CO2/kWhH2); with partial oxidation of
heavy oil this rises to even 400 g CO2/kWhH2 (Dreier and Wagner, 2001). This is
clearly more than is created through the direct use of natural gas and crude oil. If
hydrogen is extracted by electrolysis using average electric current in countries like
Britain or Germany, the carbon dioxide figure rises to between 500 and 600 g CO2/
kWhH2. Hydrogen as an energy source then ends up being far from ideal for the
environment. For the climate balance sheet it ultimately makes more sense to con-
tinue driving petrol and natural gas-powered cars than to change over to those run
on hydrogen.
Hydrogen only offers a true alternative if it is extracted using pure renewable energy
sources, such as through electrolysis using energy from wind turbines and solar
power plants. However, as long as hydrogen is produced using methods that can
create carbon dioxide, it will at best be suitable for testing prototypes.
Many manufacturers are developing products that allegedly rely on clean hydrogen
as an energy source and fuel cells for generating electricity. They owe us an expla-
nation of how they intend to make sufficiently large quantities of reasonably priced
carbon-free hydrogen available soon.
In the foreseeable future renewable energies will be able to compete fully with fossil
and nuclear energy plants on the basis of open competition. As a result, the produc-
tion costs for hydrogen produced with renewable electricity will drop. This will then
enable hydrogen and fuel cells to become an interesting component of sustainable
energy supply. Therefore, it is already a good idea to encourage the development
of these technologies.

13.5 Markets, Outlook and Development Potential

Hydrogen production is currently at a total of 30 million tons worldwide. In


comparison, crude oil consumption at 3953 million tons worldwide in 2007 was
higher by several orders of magnitude. As the chemical industry uses the bulk
of the available hydrogen, a market for it does not yet exist in the energy sector.
The capacity for hydrogen production using renewable energies is very low, and
there is also no infrastructure for the transport and storage of large volumes of
hydrogen.
Only a small number of hydrogen refuelling stations currently exist to supply
hydrogen for filling the tanks of hydrogen test vehicles (Figure 13.9). The cost of
developing a comprehensive network of stations that could provide hydrogen fuel
is estimated in the billions of euros in countries like Germany and Britain alone. In
larger countries such as the USA this sum would be even higher. As long as hydro-
gen is still much more expensive than conventional fuels, the chances of this kind
of investment are very low.

274
13.5 Markets, Outlook and Development Potential

Figure 13.9 Many car makers and energy companies are banking on hydrogen. However, the network
of hydrogen fuel stations is still extremely small and is only capable of filling up a few prototypes that
use hydrogen. Photos: BP, www.bp.com.

Whereas electric drives with batteries charged using renewable energies offer an
alternative for cars, climate-neutral concepts are still lacking for the powering of
aeroplanes. Hydrogen could be the solution in this area. However, as planes operat-
ing on hydrogen also emit water vapour and produce condensation trails, which in
turn contributes to the greenhouse effect, air traffic based on hydrogen would not
be totally climate neutral.
Even if hydrogen remains a very interesting energy source, it will take at least 10
to 20 years before it can be used on a widespread scale.
A large number of fuel cell projects today are therefore relying on natural gas as an
energy source. Even technical developments in fuel cell technology are progressing
at a much slower pace than the product announcements in the 1990s may have led
us to believe. A few commercial projects exist at the moment, but the number of
units sold is still relatively low. Technical advances are also needed in the fuel cell
area before a larger market can be exploited. Above all, a functioning renewable
hydrogen industry is needed before fuel cells can become a truly climate-compatible
alternative.

275
14
14 Sunny Prospects – Examples
of Sustainable Energy Supply
The previous chapters covered the large spectrum of renewable energies available.
Renewable energies will largely have to replace conventional energy sources during
the next 50 years to keep the consequences of climate change within manageable
limits. The modern use of these energies does not have years of tradition to fall back
upon. Their applications are often new and unusual.
Yet climate-compatible energy supply is not pie in the sky. A life beyond fossil and
nuclear energy sources is already possible today. This chapter shows the truth of
this claim based on examples, some of them quite impressive, from a variety of very
different areas.

14.1 Climate-Compatible Living

Despite the pressing need to protect the climate, low-energy dwellings such as three-
litre houses and passive houses are still the exception rather than the rule. Most
buildings that undergo restoration or renovation also fall short of meeting the stand-
ard for climate compatibility.
It is technically as well as economically easy to incorporate energy supplies that
are low or neutral in carbon dioxide into new buildings in particular. It is far
simpler to do this at the building stage rather than adapting the finished house in
the future. Energy-saving adaptations are usually postponed so long that windows,
roofs and outside walls are already due for refurbishment anyway. However,
with well constructed new buildings this can take many years. If energy prices
continue to rise over the next few years, then even new-builds with unnecessarily
high requirements for energy will become very costly to run. Investing in high-
quality insulation and renewable energy supply when a house is being built or

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

276
14.1 Climate-Compatible Living

renovated means not having to worry about increases in energy prices in the years
to come.

14.1.1 Carbon-Neutral Standard Prefabricated Houses


Most of the climate-protecting measures available in building construction are rela-
tively unspectacular. The first example shows an average single-family home in
Berlin with about 150 square metres of living space (Figure 14.1). It was built in
2005 using wood frame construction. Compared to solid construction, wood frame
construction offers good insulation values even with thinner walls. With a house of
this kind, 8 cm of additional outer insulation in a total of 32 cm reduce heat loss
through the walls. The standard thermal windows were replaced using triple-glazed
windows with a high insulation value. A controlled ventilation system with heat
recovery reduces ventilation losses and increases comfort levels. These measures
alone will enable this house to reach the standards of a three-litre house.

Figure 14.1 Left: Family home in Berlin with carbon-neutral energy supply. Right: Heating cellar with
buffer storage tank, solar cycle pump, controlled heat recovery system and wood pellet boiler.

A wood pellet boiler and a 4.8 square metre solar thermal system cover remaining
heating needs and supply hot water. A pellet store 5 square metres in size stores
about 6 tons of wood pellets. This covers heat energy requirements for three years
and is even enough to bridge periods when prices for fuels are particularly high. A
7.4 square metre photovoltaic system with output of one kilowatt peak feeds its
power into the public grid.
The additional costs for the measures described compared to the minimum new-
build standard in 2005 amounted to around 30 000 euros. The costs of wood pellets
and electricity for the ventilation motors and heating pump are around 400 euros
per year. The energy credits for feed-in solar power are roughly equal to this, so
this house actually has no additional heating costs.

277
14 Sunny Prospects – Examples of Sustainable Energy Supply

The only other energy costs are those for normal electricity consumption. A green
energy supplier offers carbon-free electricity. The photovoltaic system saves even
more carbon dioxide than is emitted during wood pellet production and the transport
of the pellets. As a result, the energy supplied to this house is completely carbon
neutral.

14.1.2 Plus-Energy Solar House


If the entire roof of a house with optimal insulation is equipped with solar panels,
this can change a carbon-neutral house into a plus-energy solar house. An example
is the house shown in Figure 14.2. Optimal heat insulation using cellulose, triple-
glass solar glazing, a ventilation system with heat recovery and preheated air sup-
plied from the ground reduce the heat energy needs by about 80% compared to a
standard new-build.

Figure 14.2 Plus-energy solar house in Fellbach. The photovoltaic system feeds more energy into the
public grid each year than the house needs for its own heating and electricity. Source: Reinhard Malz,
www.fellbach-solar.de.

A brine-water heat pump with 1.1 kilowatts of electric power covers the remaining
heat energy requirements of the house. The low temperature heat comes from two
vertical earth probes sunk 40 metres into the ground. A 0.50-kilowatt heat pump
would have been sufficient but is not available on the market.
The photovoltaic system with a power output of 8 kilowatts peak feeds at least 8000
kilowatt hours into the public grid. The volume of energy exceeds the energy
requirement of the house, including the ventilation system and the heat pump. As
a result, the energy supplied in this house is not only carbon neutral but also saves
carbon dioxide in other areas due to the feed-in solar energy.

278
14.1 Climate-Compatible Living

14.1.3 Plus-Energy Housing Estate


Houses that produce an energy surplus over the year and feed it into the grid do not
have to be an exception. This is shown by the Schlierberg solar housing estate in
Freiburg in Southern Germany. The architect Rolf Disch built 50 plus-energy houses
on this estate. He has been integrating ecological enhancements and renewable
energies into his projects for more than 30 years.
All houses in this development of terraced houses are orientated towards the south.
The gaps between the terraces are designed so that the heat from the sun can stream
into the houses through large south-facing windows in the winter. Balconies that jut
out prevent the houses from heating up too much in the summer. These simple and
cheap measures could reduce the energy needs of many new-build houses.
Optimal insulation that exceeds the standard and controlled living space ventilation
with heat recovery are standard features of this particular housing estate. These fea-
tures are designed to keep heat energy needs to an absolute minimum (Figure 14.3).

Figure 14.3 Schlierberg solar housing estate in Freiburg, Germany, with 50 plus-energy houses.
Source: Architekturbüro Rolf Disch, www.solarsiedlung.de.

A wood chip heat power plant covers the rest of the heat energy requirements and
provides climate-neutral energy supply. A district heat grid transports the heat to the
individual houses. The photovoltaic systems feed into the public electricity grid. The
solar energy surplus produced by the houses makes them plus-energy homes.
The building materials were also chosen with ecology and sustainability in mind.
The wood has not been chemically treated, solvent-free paints and varnishes have
been used throughout and the water and electric lines are PVC-free.

14.1.4 Heating Only with the Sun


Many houses use solar thermal systems only to heat water in the summer and at
best to provide heating support during the transitional seasons. It is hard to imagine

279
14 Sunny Prospects – Examples of Sustainable Energy Supply

that houses in countries with cold winters could be heated with solar energy alone.
But in recent years many single-family houses have proven that heating require-
ments can be covered by the sun even in Central Europe. The first multiple-family
dwelling in Europe to be heated 100% with solar thermal heat was built in Burgdorf
near Bern, Switzerland, in 2007 (Figure 14.4).

Figure 14.4 Left: First multiple-family dwelling in Europe to be heated 100% with solar thermal heat.
Right: Installation of 276 m2 roof-integrated solar collector. Source: Jenni Energietechnik AG, www.jenni.ch.

High-quality heat insulation and a ventilation system with heat recovery ensure that
heat losses in this building are relatively low. The large 276 square metre roof is
completely covered with solar collectors. An enormous hot water storage tank con-
taining 205 000 litres is located in the centre of the building. The seasonal storage
tank extends from the cellar to the loft. The collectors heat the storage tank with
solar heat in the summer. In winter when the collectors cannot supply sufficient
heat, the heat from the storage keeps the building toasty warm. Supplemental heat
is not required.
Less than 10% of the total construction cost of around 1.8 million euros for the
building was spent on the solar system. Yet aside from the small expenditure for
electricity for the solar and heating cycle pumps, no additional heating costs are
actually incurred. Heating costs are included in the rent, so that residents of the
block do not have to worry about constant rises in energy bills.

14.1.5 Zero Heating Costs after Redevelopment


Practically all technical options can be incorporated into new-builds without a
problem. The situation is considerably more complex with existing buildings that
date from a time before protecting the environment was considered important.
Building renovations or the installation of a new heating system present the ideal
opportunity to make enormous reductions in carbon dioxide. One example is the

280
14.2 Working and Producing in Compatibility with the Climate

zero-heating cost apartment block in Ludwigshafen, Germany, which was built in


the 1970s and then renovated in 2007 to a technical standard comparable to today’s
levels (Figure 14.5).

Figure 14.5 Constant hikes in energy bills are a thing of the past for this zero-heating cost house in
Ludwigshafen, Germany. Source: LUWOGE, www.luwoge.de.

The result is a reduced heat energy requirement of around 80%, which now amounts
to only about 20 kilowatt hours per square metre of living space per year. Thirty-
centimetre thick outer wall insulation and triple-glazed thermal windows provide
optimal building insulation. Controlled living space ventilation with heat recovery
reduces ventilation heat losses. Solar thermal façade collectors provide for the hot
water supply and a large grid-coupled photovoltaic system is located on the roof.
The minimal remaining heating needs are covered electrically. The yield from the
photovoltaic system should be roughly the same as the heating electricity costs. The
tenants do not pay for any heating. Constant rises in heating costs are therefore also
a thing of the past for this building. Taking into account all costs including upkeep
and vacancy rates, the modernization measures have also turned out to be very
lucrative for the investor. Both sides benefited from the climate-protection measures
that were taken.

14.2 Working and Producing in Compatibility


with the Climate

14.2.1 Offices and Shops in Solar Ship


The Sonnenschiff (Solar Ship) is located near the housing estate in Freiburg
discussed above. The Sonnenschiff is a solar service centre that houses shops,

281
14 Sunny Prospects – Examples of Sustainable Energy Supply

offices, medical practices and penthouse flats. The architect Rolf Disch also planned
this complex as a plus-energy construction. Optimal insulation incorporating modern
vacuum processes and intelligent ventilation with heat recovery help reduce heat
requirements. In winter the sun streams through large windows right into the heart
of the building, providing some of the heating. In summer canopies and blinds
prevent the building from overheating. A wood chip heat power plant covers the
minimal remaining heating needs. Photovoltaic systems are not simply mounted on
the roof but are actually used instead of a conventional roof. As a result, the entire
roof surface generates electric power (Figure 14.6).

Figure 14.6 The Sonnenschiff or solar ship contains two shops, a bistro, offices and medical practices.
Source: Architekturbüro Rolf Disch, www.solarsiedlung.de.

An intelligent solution has also been devised for transport connections. The complex
has its own tram stop and conveniently sited parking areas for bicycles, thereby
offering alternatives to motor car use. The electric cars of a car-sharing firm
are allocated preferred parking spaces and can be charged using solar power gener-
ated by the complex itself. The Sonnenschiff complex has put into practice the
concept of working and living carbon-free and contributing to climate-compatible
mobility.

14.2.2 Zero-Emissions Factory


It is also relatively easy for factories to operate without emitting carbon dioxide.
An example is the Solvis zero-emissions factory in Braunschweig, Germany, which
produces solar and pellet systems. High-quality heat insulation and waste air recov-
ery in the factory have resulted in energy savings as high as 70% compared to
conventional structures. Optimal use of daylight and low-energy office machines
have halved electricity needs. Solar collectors and photovoltaic systems generate
30% of the electricity and heating required (Figure 14.7). A rapeseed oil block-
heating plant covers remaining energy needs, thereby ensuring that all energy supply
is completely carbon-free.

282
14.2 Working and Producing in Compatibility with the Climate

Large amounts of heat often escape through the open shed doors of factories
during the loading and unloading of heavy goods vehicles. This factory keeps heat
losses at a minimum because the loading and unloading zones are situated inside
the building. A direct connection to local public transport, parking facilities for
bicycles and showers for cyclists provide employees with a climate-compatible
workplace.

Figure 14.7 Solvis zero-emissions factory in Braunschweig, Germany. Source: SOLVIS, www.solvis.de.

14.2.3 Carbon-Free Heavy Equipment Factory


The ‘factory of the future’, completed in May 2000 by Wasserkraft Volk AG, is the
first energy self-sufficient carbon-free heavy equipment factory built in Germany.
The administration building is mainly built of timber from the Black Forest
area. The entire building including the factory floor is very well insulated. The
offices are orientated towards the south, and the factory has generous overhead
lighting.
This construction enables maximum use of daylight and the incoming sunlight
covers some of the heating needs. A 30 square metre sun collector on the adminis-
tration building actively uses solar energy. At the heart of the factory’s energy
supply is its own hydropower plant with 320 kilowatts of output (Figure 14.8). The
turbines integrated into the building use natural water from the nearby Elz River.
The waste heat from the generator covers about 10% of the heat requirement. Three
heat pumps with a heat output of 130 kilowatts extract heat from the groundwater
and supply the remaining heat energy.
The hydropower plant generates an annual surplus of around 900 000 kilowatt hours
of electric energy, which the factory feeds into the public grid. This makes it not
only a carbon-free factory but also a plus-energy one that supplies additional
ecological electricity.

283
14 Sunny Prospects – Examples of Sustainable Energy Supply

Figure 14.8 The administration building at the Wasserkraft Volk AG factory of the future in Gutach,
Germany, has its own hydropower plant to generate carbon-free power. Source: WKV AG, www.wkv-ag.de.

14.3 Climate-Compatible Driving

The colourful, glossy brochures produced by leading car manufacturers are full of
impressive assertions about how much they are doing to protect the environment.
However, with many manufacturers it is difficult to discern any real innovation in
creating carbon-free means of transport. Cars that run on natural gas and hybrid cars
make only a marginal difference in reducing carbon dioxide emissions if one looks
at the recommendations for reductions set as a standard for protecting the climate.
The sustainable cultivation potential with biomass fuels is too low to meet total
needs and there is simply not enough carbon-free hydrogen being produced to
advance the innovation of hydrogen automobiles.
Plug-in hybrid vehicles or pure electric cars could offer an alternative in the short
term. Plug-in hybrid vehicles are normally refuelled through a power point and for
emergencies have a combustion engine that can also be run on biofuel. Pure electric
cars only run on electricity. If the electricity is supplied by renewable power plants,
no carbon dioxide is emitted while the vehicle is being driven.

14.3.1 Waste Gas-Free Electropower


Two examples from the USA confirm that electric cars do not have to be dull and
utilitarian. The companies Tesla Motors and Commuter Cars are currently develop-
ing really innovative vehicles (Figure 14.9).
Tesla Motors offers a 185 kilowatt electric engine for acceleration from 0 to 100 km/h
in less than four seconds. A modern lithium-ion battery enables a range of 390 km
on one battery charge. The battery service life should last up to 160 000 km. It takes
three and a half hours to charge the battery.

284
14.3 Climate-Compatible Driving

Figure 14.9 Modern electric cars are now far from boring. Photo left: Tesla Motors, www.teslamotors.com.
Right: Commuter Cars, www.commutercars.com.

Whereas a combustion engine achieves an average efficiency of less than 30%, an


electric engine reaches an impressive 90%. The consumption is therefore only
around 13 kilowatt hours per 100 km, which converts to 1.3 litres of petrol per
100 km. Even with electricity from fossil power plants the carbon dioxide emissions
of this electric sports car are below those of a fuel-efficient small car with a com-
bustion engine. If the electricity used for refuelling is sourced from a green electric-
ity supplier, this type of car can be driven without emitting any carbon dioxide
whatsoever.
The Tango car from Commuter Cars also accelerates from 0 to 100 in four seconds.
The range is between 100 and 250 km. The charging time is normally three hours
but a fast charge option is available that takes ten minutes to recharge for an 80 km
distance.
Both vehicles are currently only produced in very small numbers, and they cost
around 70 000 euros. But these examples show that climate-compatible alternatives
do exist. If the large car manufacturers made a serious effort to develop these alter-
natives, climate-neutral driving would finally become a reality.

14.3.2 Travelling around the World in a Solar Mobile


Solar cars have a reputation of being at best suitable for driving to the postbox and
back. They are usually not trusted to be driven for long distances under extreme
conditions. To highlight the problems of climate change, a Swiss man, Louis Palmer,
set off on the first trip round the world in a solar-operated taxi in July 2007. Palmer
had been dreaming about the idea of a solar taxi since 1986. In 2005 he finally suc-
ceeded in implementing the project, known as ‘Taxi’, thanks to the help of numerous
sponsors and technical support from the Swiss Technical University in Zurich
(Figure 14.10).

285
14 Sunny Prospects – Examples of Sustainable Energy Supply

Figure 14.10 In July 2007 Louis Palmer of Switzerland began his trip around the world in a solar
energy-operated vehicle. Photos: www.solartaxi.com.

His route took him over 50 000 km through 50 countries on five different continents
in 15 months. On the trip he and his team took advantage of numerous stops and
events to introduce solar techniques and provide impetus for the use of new climate-
friendly technologies. After taking a test drive many interested people in the coun-
tries visited became enthusiastic about the technology.
The solar taxi is designed as an electric car. It has a 5-metre trailer that is surfaced
with six square metres of solar cells. A new type of battery stores the electricity,
thereby enabling the car to be driven at night and when there is no sun. The daily
range is about 100 km. If long distances are to be driven, the battery is recharged
using additional solar energy from the grid. A photovoltaic system was installed
specifically for this purpose on a roof in Bern, Switzerland.

14.3.3 Across Australia in Thirty-Three Hours


The World Solar Challenge is proving that solar cars are already capable of extremely
high performance. This event has been taking place regularly in Australia since 1987
and is the most rigorous race in the world for solar cars. The race follows a route
on public roads for around 3000 km right across Australia – from Darwin in the
north to Adelaide on the south coast. The racing teams try to cover as great a distance
as possible between the hours of eight in the morning and five in the afternoon.
The vehicles battling for first place in the race are technical masterpieces. The size
of the batteries is restricted by the rules and the only energy allowed is what is
supplied by solar modules that are mounted directly onto the vehicles. Powerful
solar cells with efficiencies of well over 20% provide the necessary drive energy.
The cars are optimized for aerodynamics and trimmed to a minimal weight. As a
result of constant technical advances and more and more efficient solar cells, the
average speeds of the winners have been increasing steadily. In 2005 it was

286
14.3 Climate-Compatible Driving

Figure 14.11 In 2007 the solar car from Bochum University reached average speeds of 73 km/h during
the 3000 km race. Photos: Bochum University of Applied Sciences, http://www.hs-bochum.de/solarcar.html.

102.7 km/h. As Australia imposes speed limits on public roads, it will not be pos-
sible from a practical standpoint to increase this speed much further. Therefore, in
2007 the rules were changed to include limiting the size of the solar generator to
six square metres and stipulating a seated position for the driver.
Teams from all over the world have been participating in the competition for many
years. In 2007 the Solarworld No. 1 racing car from Bochum University of Applied
Sciences reached fourth place (Figure 14.11). It took the team from Bochum around
41 hours of pure driving time to cover the 3000 km distance. The winning car Nuon
Nuna II from the Netherlands just needed 33 hours.

14.3.4 Game over CO2!


The German solar company SOLON AG presented a 100% carbon-free mobility
concept in 2007 with the slogan ‘Game over CO2!’. The central element of
this concept is a new electric motorbike developed by the Vectrix company in
the USA.
The two-seater reaches a peak speed of 100 km/h. Its good acceleration behaviour
enables it to move quickly, particularly in city traffic. The nickel-metal-hybrid
battery provides a range of 55 to 90 km of driving and can be recharged in two
hours.
At the filling station of the future, which already operates successfully in a number
of cities, tracked photovoltaic modules are used to charge the electric motorbikes
(Figure 14.12). In the meantime the electric bike is being produced in unit quantities.
Even if purchasers do not have a solar filling station nearby, they can operate
the motorbike carbon-free using a photovoltaic system on a house roof or with
renewable electricity from a green energy supplier. The Internet page www.
game-over-co2.de makes very interesting reading, even for non-bikers.

287
14 Sunny Prospects – Examples of Sustainable Energy Supply

Figure 14.12 The filling station of the future was launched in Austria in 2007. It refuels electric
motorbikes with carbon-free solar power. Photos: SOLON AG for solar technology. Photographer of photo left:
www.marcusbredt.de.

14.4 Climate-Compatible Travel by Water or Air

14.4.1 Modern Shipping


As a result of globalization, goods are being shipped over ever-longer distances. A
large part of the increase in transport is being handled by commercial shipping. The
carbon dioxide emissions per transport kilometre are significantly lower with ship-
ping compared to air freight. Nevertheless, shipping is also contributing noticeably
to the greenhouse effect. Two to three percent of all global carbon dioxide emissions
are attributed to shipping – and the rate is rising.
Until the middle of the nineteenth century, sailing ships dominated freight and pas-
senger traffic at sea. Steamships then came along and had the advantage that they
did not have to depend on wind conditions to keep to their timetables. They gradu-
ally replaced sailing craft, which today are used almost exclusively for leisure and
sporting activities.
Yet new types of concepts exist that make wind power useable in combination with
conventional ship propulsion. The German inventor Anton Flettner developed a
cylindrical rotor to propel ships in the 1920s. However, this drive did not catch on
at the time. New ship prototypes are currently in development, including the Flettner
rotor combined with a conventional ship diesel drive, aimed at reducing fuel require-
ments from 30 to 40%.

288
14.4 Climate-Compatible Travel by Water or Air

Another interesting concept is based on modern power kites (Figure 14.13). A start-
ing and landing system automatically lowers and raises the power kite and the
cabling rope. The kite flies at a height between 100 and 300 m. The wind conditions
at this height are higher and more constant than on the deck of a conventional sailing
ship. The cabling rope, which is made of modern synthetic material, is attached to
the forepart of the ship and transfers the driving force there to the ship. The steering
of the kite is completely automated. Through a shortening or lengthening of the
steering lines, a power kite can be controlled like a paraglider and orientated with
optimal precision depending on wind direction, wind intensity and a ship’s course.
With a sail surface of up to 5000 square metres, the power output should reach up
to 5000 kilowatts or 6800 HP. On a yearly average a power kite should be able to
reduce fuel requirements, and consequently carbon dioxide emissions, by 10 to 35%.
Under optimal conditions savings of up to 50% are sometimes possible.
Modern wind drives can therefore make a considerable contribution towards reduc-
ing carbon dioxide emissions. Emissions in shipping can even be eliminated com-
pletely if biofuel or renewably produced hydrogen is used to cover the rest of the
fuel requirements of conventional ship drives.

Figure 14.13 Modern automatically controlled power kites can reduce the fuel requirements of
conventional shipping by up to 50% and, when combined with biofuels, even make it carbon-free.
Photos © SkySails, www.skysails.de.

14.4.2 Solar Ferry on Lake Constance


Solar boats offer climate-compatible travel on water even if there is no wind, espe-
cially for short distances. The Helio solar ferry on Lake Constance has been connect-
ing the German town of Gaienhofen to Steckborn in Switzerland since May 2000
(Figure 14.14). The 20 m ship can carry 50 people, and electric engines with an output
of 8 kilowatts each provide for a maximum speed of 12 km/h. The batteries of the
boat ensure a range of 60 to 100 km of travelling distance. The roof of the boat consists

289
14 Sunny Prospects – Examples of Sustainable Energy Supply

of an optically very successful photovoltaic system that produces an output of 4.2


kilowatts, which provides most of the power needed for charging the batteries.
In addition to its climate-compatibility, this solar boat also offers other environmen-
tal benefits. It runs very quietly and, unlike conventional diesel engine ships, does
not emit unpleasant exhaust fumes. The construction of the ship only causes minimal
waves and, as a result, does not contribute towards any further erosion on the lake’s
shores.

Figure 14.14 The Helio solar ferry has been crossing Lake Constance since 2000. Photos: Bodensee-
Solarschifffahrt, www.solarfaehre.de.

14.4.3 World Altitude Record with a Solar Aeroplane


Hot-air balloons were the first flying machines people used. Fire from firewood or
straw produced the hot air needed for the carbon-neutral powering of a balloon.
Today hot-air balloons are usually powered by natural gas burners. However, these
balloons are highly unsuitable for freight transport or regular services. Without
exception, propeller and jet-powered airplanes rule commercial aviation. The kero-
sene used is produced from oil, so the prospect of climate-friendly air travel is still
a long way off.
But an unmanned light aeroplane called Helios, after the Greek sun god, shows that
fossil fuels and flying do not have to be inextricably linked (Figure 14.15). The
plane was developed by NASA and the California company AeroVironment and
had its maiden flight in 1999. A total of 62 130 silicon solar cells with an efficiency
of 19% are located on the wings, which have a span of 75.3 m and a depth of 2.4 m.
These solar cells deliver the energy for 14 electric engines with a total output of 21
kilowatts. Powerful lithium batteries enable the plane to fly even after sunset.
Due to the low power output, the flight speed at low altitudes did not even reach
45 km/h. However, the performance of this aeroplane is not attributed to its speed
but to its flying altitude. Flying over Hawaii at an altitude of 29 524 m on 13 August

290
14.4 Climate-Compatible Travel by Water or Air

2001, this plane set a world record for non-rocket-operated planes. Sadly, on 29
May 2003 it broke apart during a test flight and plunged into the Pacific Ocean near
Hawaii.

Figure 14.15 NASA’s Helios aeroplane, powered only by solar cells, set a new world record for
altitude in 2001. Photos: NASA, www.dfrc.nasa.gov.

14.4.4 Flying around the World in a Solar Plane


The Swiss psychiatrist, scientist and adventurer Bertrand Piccard is hoping to have
more luck with his solar aeroplane. He is mainly known for his trip around the world
in a hot-air balloon in 1999. His newest project is called Solar Impulse. What he
wants to do is circle the world in a glider that is powered only by solar energy. This
undertaking is designed to be an effective PR exercise in proving the technical pos-
sibilities offered by renewable energies. Unlike the NASA Helios aeroplane, this
will be a manned flight (Figure 14.16).
The planned start date is sometime in 2011. The first prototype of the aeroplane has
already been built. It has a wingspan of around 61 m, weighs 1500 kg and is designed
to fly at an average speed of 70 km/h. The main considerations in the construction
of the plane were the incorporation of optimal aerodynamic properties and extreme
lightness in weight. The wing surfaces are totally covered with solar cells. These
supply enough energy to power the aeroplane without the need for any additional
energy. During the day high-performance batteries store some of the solar energy
so that the plane can also fly at night.
Whereas the prototype is only constructed for a 36-hour flight, a second aeroplane
is being developed with enhanced capabilities to enable several days of flying time
for flights around the world. This is necessary for crossing the Atlantic and bridging
other large distances. During the daytime the solar power propels the aeroplane to
heights up to 12 000 m. Powered by the solar energy stored in the batteries, the plane
will then be able to hold steady at an altitude of around 3000 m until the morning.

291
14 Sunny Prospects – Examples of Sustainable Energy Supply

Five stopovers are planned during the flight around the world. Each leg of the
flight should take four to five days, which is as much as can be expected from one
pilot. During stopovers the pilot will be changed and promotional activities will be
organized.
With technical advances in batteries and a resulting anticipated reduction in weight,
the aeroplane could be used for longer flights with two pilots. This would then
bring a non-stop trip around the world in a solar aeroplane into the realm of
possibility.

Figure 14.16 Bertrand Piccard of Switzerland aims to fly the aeroplane Solar Impulse round the world
in 2011. Photos © Solar Impulse/EPFL Claudio Leonardi.

14.4.5 Flying for Solar Kitchens


Based on current technology, it does not seem likely that solar aeroplanes will ever
be able to replace large conventional planes. Even with highly efficient solar cells,
the space available on the surface of the wings is not large enough to provide suf-
ficient driving energy for planes carrying loads of several hundred tons. With the
exception of a very limited possible use of biofuels, there are no options offering
complete climate-compatible air travel. In the long term renewably produced hydro-
gen offers an alternative to fossil fuels.
Until then the only carbon-free alternative available is no air travel at all. Modern
communication technologies and more and more interesting leisure activities in
close proximity to where people live are helpful alternative options to travel. But
the solarium around the corner is not really a substitute for a winter break in a sunny
resort. It makes it even more difficult to stick to one’s principles if cheap flights are
offered at the same price as a ticket for the local city train.
Section 3.5 of this book showed that, as an interim solution, investments in other
areas can reduce as much carbon dioxide as created by an unavoidable flight. The
carbon dioxide emissions of a flight from Berlin to New York and back can be
offset for around 100 euros. As part of its programme, the non-profit company
Atmosfair offers measures designed to compensate for emissions and at the

292
14.5 Carbon-Free Electricity for an Island

same time recommends considering videoconferencing and travel by train as


alternatives.
For example, Atmosfair had large solar thermal systems installed in temples, hos-
pitals and schools in different projects at 18 locations in India. One of the projects
is a solar kitchen for a Hindu place of pilgrimage, Sringeri Mutt. Diesel burners had
been supplying the energy to prepare meals for thousands of pilgrims until modern
solar mirrors replaced them as part of the project (Figure 14.17). The mirrors bundle
sunlight onto a pipe and heat the water, which is then fed to the kitchen. A cleverly
devised steam system ensures that the kitchens can still function even after sunset.
Another aim of the project is to implement a transfer of technology to local enter-
prises. The systems for all 18 projects were made by an Indian manufacturer and
by 2012 should be offsetting a total of around 4000 tons of carbon dioxide emissions
from air travel.

Figure 14.17 Solar-mirror systems financed through contributions from environmental protection funds
replace conventional diesel burners in large kitchens in India. Source: atmosfair, www.atmosfair.com.

14.5 Carbon-Free Electricity for an Island

The combination power plant project described in Chapter 4 showed that completely
carbon-free and reliable electricity supply based on renewable energies is techni-
cally possible if different renewable plants work together on a complementary basis.
The prerequisite for renewable electricity supply is a large-scale grid interconnec-
tion, which is available almost everywhere in densely populated industrialized
countries. Smaller grids in remote regions can also guarantee supply using indi-
vidual renewable electricity power plants combined with storage facilities.
One example is the Utsira community in Norway. Around 215 inhabitants live here
on a 6 km square island 18 km west of the Norwegian mainland. In a pilot project

293
14 Sunny Prospects – Examples of Sustainable Energy Supply

started in 2004 two 600-kilowatt wind turbines are providing ten of the households
with their complete supply of electricity (Figure 14.18). The electricity demand in
Norwegian households is generally rather high as it is also often used to provide
electric heating.

Figure 14.18 Two wind turbines on the Norwegian island of Utsira guarantee a carbon-free electricity
supply for ten households. Source: Norsk Hydro ASA, www.hydro.com.

In contrast with a large interconnected grid system, this type of island grid offers
some major technical challenges. For example, if two houses switch on their electric
heating at the same time, major fluctuations will occur in the grid. Although the
region is very windy, for safety reasons the wind turbines do not deliver any elec-
tricity during the rare periods of calm or when extreme storms occur. A reliable
electricity supply also has to be available for these periods.
For this purpose a module for energy storage was tested in Utsira for future energy
supply. When surplus capacity exists, an electrolyzer converts the wind power
into hydrogen and oxygen. When the wind turbines supply too little electricity, a
60-kilowatt fuel cell makes up the difference using the stored hydrogen. In addition,
when minor fluctuations occur, a flywheel storage device provides for a stable grid.
For cost reasons the testing of the hydrogen storage technology was first limited to
two years and was then extended in 2006.

14.6 All’s Well that Ends Well

The variety of possibilities for using renewable energies described in this book, not
least the examples in this chapter, have shown that it is actually unnecessary to
endanger our future existence through an unrestrained use of crude oil, natural gas,
coal and nuclear energy. The effects of fossil energy on the world’s climate are
already noticeable. The consequences of climate change will soon be beyond our

294
14.6 All’s Well that Ends Well

control if we do not take the bull by its horns and radically restructure our energy
supply.
Renewable energies can already guarantee complete, affordable and climate-com-
patible coverage of our energy requirements. There is no basis for the widespread
fear that the lights will go out if we cannot use oil, natural gas, coal and nuclear
energy. On the contrary, increasing the use of renewable energies will make us more
and more independent of the conventional energy sources that are now in short
supply and steadily becoming more expensive. The use of renewable energies will
also help to end some of the conflicts that are taking place at the moment.
The question that remains is why we are still using renewable energies on such a
small scale. There are many reasons for this. First of all, the wider public is not
being made sufficiently aware of the possibilities offered by renewable energies. An
increase in information sources such as this book will, it is hoped, gradually close
this gap in knowledge. On the other hand, everyone is hoping that the problem will
somehow solve itself, many people do not really feel responsible or they have
already given up hope. However, it is up to each and every individual to make a
contribution and to demand the necessary measures from the politicians. If all of us
start to use the possibilities open to us to save energy and use renewable energies,
we will still be able to stop climate warming and establish a sustainable energy
supply contribution.

295
Appendix

A.1 Energy Units and Prefixes

The legal unit of energy is called a watt second (Ws) or joule (J). The most common
units of energy are normally measured in kilowatt hours (kWh). Table A.1 sum-
marizes the units commonly used in the energy sector. Table A.2 shows the elements
and prefixes for energy units.

Table A.1 Conversion factors between different units of energy.

kJ kcal kWh kg coal kg oil m3 natural


equivalent equivalent gas
1 kilojoule 1 0.2388 0.000 278 0.000 034 0.000 024 0.000 032
(1 kJ = 1000 Ws)
1 kilocalorie (kcal) 4.1868 1 0.001 163 0.000 143 0.0001 0.000 13
1 kilowatt hour (kWh) 3600 860 1 0.123 0.086 0.113
1 kg coal equivalent 29.308 7000 8.14 1 0.7 0.923
1 kg oil equivalent (oe) 41.868 10 000 11.63 1.428 1 1.319
1 m3 natural gas 31.736 7580 8.816 1.083 0.758 1

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

296
A.2 Geographic Coordinates of Energy Power Plants

Table A.2 Elements and prefixes.

Prefix Abbrev. Value Prefix Abbrev. Value


Kilo k 3
10 (thousand) Milli M 10−3 (thousandth)
Mega M 106 (million) Mikro μ 10−6 (millionth)
Giga G 109 (billion) Nano N 10−9 (billionth)
Tera T 12
10 (trillion) Piko P 10−12 (trillionth)
Peta P 1015 (quadrillion) Femto F 10−15 (quadrillionth)
Exa E 1018 (quintillion) Atto A 10−18 (quintillionth)

A.2 Geographic Coordinates of Energy Power Plants

Most energy power plants are so large that they are easy to detect in satellite images.
The free software Google Earth enables satellite images with a high resolution to
be viewed on one’s own computer at home. The following list shows the coordinates
of some interesting systems and power plants. The coordinates can be entered
directly into Google Earth but must comply with the following format:
+51° 23′ 23′′ , +30° 05 ′ 58′′
This stands for 51°23′23″ N and 30°05′58″ E. Negative prefixes are used with S
and W.

■ earth.google.co.uk Google Earth start page and free download

Conventional Energy Power Plants


51°23′23″ N Shut down nuclear energy plant in Chernobyl with 4 blocks
30°05′58″ O each with 1000 MW. A reactor accident occurred in Block IV
on 26.4.1986. Today a cement sarcophagus encloses the
damaged reactor.
51°45′47″ N Fast breeders at Kalkar; built for 3.6 milliard euros and never
6°19′44″ O put into operation. Today the Wunderland Kalkar leisure park
is situated on the grounds of the former nuclear energy plant.
51°50′00″ N Jänschwalde brown coal-fired power plant containing 6 blocks
14°27′30″ O each with 500 MW. Built between 1976 and 1989. With an
output of 25 million tons of carbon dioxide per year, this power
plant has the highest emissions of any plant in Germany.
51°04′30″ N Garzweiler II brown coal opencast mine. Around 1.3 tons of
6°27′00″ O brown coal are to be extracted from an area of 48 square
kilometres by 2045. As a result, 12 towns with 7600
inhabitants are to be relocated.

297
Appendix

Photovoltaic Systems
48°08′08″ N 2.7-MW roof PV system with 21 900 PV modules on the new
11°41′55″ O Munich exhibition centre. Built in 1997, expanded in 2002 and
2004.
39°49′55″ N 1-MW open space PV system with 7936 PV modules in Toledo,
4°17′55″ W Spain. Built in 1994.

52°31′29″ N Roof-integrated PV system on the main train station in Berlin.


13°22′05″ O 780 photovoltaic modules supply an output of 189 kWp.

Solar Thermal Power Plants


37°05′42″ N Plataforma Solar de Almería European test centre in Spain
2°21′40″ W with solar tower and parabolic trough test fields.

42°29′41″ N Solar melting furnace in Odeillo (France), completed in 1970.


2°01′45″ O With 20 000-fold concentration, 63 heliostats with a total area of
2835 square metres reach temperatures of almost 4000 °C.
34°51′43″ N Solar thermal power plant at Daggett (California, USA). SEGS I
116°49′41″ W with 13.8 MW from 1985 and SEGS II with 30 MW from 1986
plus 10-MW Solar Two solar tower-test power plant from 1998.
35°00′58″ N Solar thermal power plant at Kramer Junction (California, USA).
117°33′40″ W SEGS III to SEGS VII each with 30 MW from 1987 to 1989.

35°01′56″ N Solar thermal trough power plant at Harper Lake (California,


117°20′50″ W USA). SEGS VIII and SEGS IX each with 80 MW from 1990 to
1991.
35°48′00″ N Solar thermal trough power plant Nevada Solar One (Nevada,
114°58′35″ W USA) with output of 64 MW, commissioned in 2007.

Wind Farms
55°41′32″ N Middelgrunden offshore wind farm (Denmark) with 20 wind
12°40′14″ O turbines each with 2 MW. Built in 2001.

32°13′48″ N Horse Hollow Wind Energy Centre near Abilene, Texas (USA).
100°02′50″ W 421 wind turbines supply 735 MW. Built in 2006.

Hydropower Plants
25°24′27″ S Itaipú hydropower plant (Brazil/Paraguay). 20 turbines supply a
54°35′19″ W total output of 14 GW. The dam wall is 7760 m long and 196 m
high.

298
A.2 Geographic Coordinates of Energy Power Plants

30°49′10″ N Three Gorges Dam in China. Built between 1993 and 2006. 26
111°00′00″ O turbines supply output of 18 200 MW.

36°00′58″ N Hoover Dam (Nevada-Arizona, USA). Built between 1931 and


114°44′16″ W 1935. Height of dam wall 221 m. Total of 17 turbines with an
output of 2074 MW.
47°33′22″ N Run-of-river power plant Laufenburg on the Rhine in Germany.
8°02′56″ O Completed 1914. Electricity output 106 MW.

47°34′12″ N Rheinfelden run-of-river power plant on the Rhine in Germany.


7°48′46″ O First run-of-river plant in Europe, built in 1899. Currently being
modernized and enlarged.
50°30′34″ N Pumped storage power plant Goldisthal in Thüringen, Germany.
11°01′16″ O The upper basin contains 12 million m3 of water. The power
plant output is 1060 MW.
48°37′08″ N Rance tidal power plant (Saint-Malo, France). Completed in
2°01′11″ W 1966. 24 turbines supply 240 MW.

299
References

Arrhenius, S. (1896) On the influence of carbonic acid in the air upon the temperature of
the ground. The London, Edinburgh and Dublin Philosophical Magazine and Journal of
Science, 5, 237–276.
Bard, J., Caselitz, P., Giebhardt, J. and Peter, M. (2004) Erste meeresströmungsturbinen-
pilotanlage vor der englischen küste. Tagungsband Kassler Symposium
Energie-Systemtechnik.
Bundesverband Solarwirtschaft (ed.) (2009) Statistische Zahlen der deutschen Solarwirt-
schaft, BWS, Berlin, www.solarwirtschaft.de.
Bundesverband WindEnergie e.V. BWE/DEWI (2008) Investitions- und Betriebskosten,
http://www.wind-energie.de/de/technik/projekte/kosten.
Deutsches Windenergie-Institut GmbH DEWI (2002) Studie zur aktuellen Kostensituation
2002 der Windenergienutzung in Deutschland, Deutsches Windenergie-Institut GmbH
DEWI, Wilhelmshaven.
Dreier, T. and Wagner, U. (2001) Perspektiven einer Wasserstoff-Energiewirtschaft. BWK,
53 (3), 47–54 und 52 (2000) Nr. 12, 41–54.
EnergieAgentur NRW (2006) Stromcheck für Haushalte, EnergieAgentur NRW,
Wuppertal, www.energieagentur.nrw.de.
Enquete-Kommission ‘Vorsorge zum Schutz der Erdatmosphäre’ des 11. Deutschen
Bundestages (ed.) (1990) Schutz der Erdatmosphäre, Economica Verlag, Bonn.
European Solar Industry Federation ESTIF (ed.) (2003) Sun in Action II, ESTIF, Brüssel,
www.estif.org.
Fichtner (2003) Die Wettbewerbsfähigkeit von großen Laufwasserkraftwerken im
liberalisierten deutschen Strommarkt. Bericht für das Bundesministerium für Wirtschaft
und Arbeit.
Fritsche, U.R. and Eberle, U. (2007) Treibhausgasemissionen durch Erzeugung und
Verarbeitung von Lebensmitteln, Öko-Institut Darmstadt.

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

300
References

Gasch, R. and Twele, J. (eds) (2004) Wind Power Plants, Earthscan, London.
German Federal Ministry for the Environment, Nature Conservation and Nuclear Safety
(ed.) (2007) Renewable Energy Sources in Figures, BMU, Berlin, http://www.erneuerbare-
energien.de/inhalt/3860/.
IEA Solar Heating and Cooling Programme SHC (ed.) (2008) Solar Heating Worldwide
2008, IEA SHC, Graz, www.iea-shc.org.
International Energy Agency, IEA (ed.) (2003) Renewables Information 2003, IEA, Paris,
www.iea.org.
International Energy Agency, IEA (ed.) (2008) Key World Energy Statistics 2008, IEA,
Paris, www.iea.org.
Intergovernmental Panel on Climate Change, IPCC (ed.) (2005) Carbon Dioxide Capture
and Storage, Cambridge University Press, New York, www.ipcc.ch.
Intergovernmental Panel on Climate Change, IPCC (ed.) (2007) Climate Change 2007:
The Physical Science Basis, IPCC, Genf, www.ipcc.ch.
Kaltschmitt, M., Merten, D., Fröhlich, N. and Moritz, N. (2003) Energiegewinnung aus
Biomasse, Externe Expertise für das WBGU-Hauptgutachten, http://www.wbgu.de/wbgu_
jg2003_ex04.pdf.
Kemfert, C. (2007) Klimawandel kostet die deutsche Volkswirtschaft Milliarden, in
Wochenbericht des (ed. KF Zimmermann), DIW, Berlin, pp. 165–169.
Kleemann, M. and Meliß, M. (1993) Regenerative Energiequellen, Springer Verlag, Berlin.
König, W. (ed.) (1999) Propyläen Technikgeschichte, Propyläen Verlag, Berlin.
Lokale Agenda-Gruppe 21 Energie/Umwelt in Lahr (ed.) (2007) Leistungsfähigkeit von
Elektrowärmepumpen, Zwischenbericht, Lahr.
Mener, G. (1998) War die Energiewende zu Beginn des 20. Jahrhunderts möglich?, in
Sonnenenergie Heft, DGS, pp. 40–43.
Messe Hamburg WindEnergy (ed.) (2006) Wind Energy Study 2006, Messe Hamburg
WindEnergy, Hamburg.
Munich Re Group (ed.) (2006) Topic Geonatural Catastrophes 2006, Munich Re Group,
Munich, www.munich-re.com.
Quaschning, V. (2007) Regenerative Energiesysteme, Hanser Verlag, München, 5.
Auflage.
Quaschning, V. (2005) Energieaufwand zur Herstellung regenerativer Energieanlagen,
http://www.volker-quaschning.de/datserv/kev.
Rahmstorf, S. (1999) Die Welt fährt Achterbahn, in Süddeutsche Zeitung 3./4.
Rahmstorf, S. and Neu, U. (2004) Klimawandel und CO2: haben die ‘Skeptiker’ recht?,
Potsdam-Institut für Klimafolgenforschung, Potsdam, www.pik-potsdam.de.
Schellschmidt, R., Hurter, S., Förster, A. and Huenges, E. (2002) Germany, in Atlas of
Geothermal Resources in Europe (eds S. Hurter und R. Haenel), Office for Official
Publications of the EU, Luxembourg, pp. 32–35.
Tscheutschler, P., Nickel, M. and Wernicke, I. (2007) Energieverbrauch in Deutschland, in
BWK Heft (ed. VDI), Springer, VDI, pp. 6–13.

301
References

Umweltbundesamt (ed.) (2007) Stromsparen ist wichtig für den Klimaschutz,


Umweltbundesamt, Berlin.
United Nations Framework Convention on Climate Change UNFCCC (ed.) (2008)
National Greenhouse Gas Inventory Data for the Period 1990–2006, UNFCCC, Posen,
www.unfccc.de.
US Department of Energy DOE (ed.) (2008) Annual Energy Review 2007, US Department
of Energy DOE, Washington, http://www.eia.doe.gov/emeu/aer.
Vattenfall Europe (ed.) (2006) The Vattenfall CO2-Free Power Plant Project, Vattenfall,
Berlin.

302
Index

absorber 117, 146 bulb turbine 195


coating 121, 148 bulk-stacked metre 243
selective 122 bypass diodes 95
swimming pools 120
absorption heat pump 227 C4 plant 239
adsorption heat pump 228 calculation
air collector 122 battery capacity 108
alkaline electrolysis 269 collector efficiency 119
annual coefficient of performance 225 collector size 135
anyway energy 16 firewood boiler output 256
Archimedes 146 heat pump COP 225
arctic ice coverage 27 hydropower plant output 205
atomic bomb 9 installable photovoltaic power103
photovoltaic efficiency 90
barrel 6 photovoltaic system yield 106
battery capacity 108 solar power plant output 157
Beaufort wind scale 170 storage size 135
Betz power factor 171 wind power 169
biofuel 270 wind power annual yield 183
biodiesel 251 wood pellet store size 257
bioethanol 252 calorific value of wood 244
biogas 254 carbon dioxide 35
biogas plant 255 balance sheet 63
biomass 237 biomass use 261
ecology 260 concentration 33
economics 259 countries of world 31
fuels 250 food stuff 64
heating 241 Germany 43
heating plant 249 hydrogen production 274
markets 263 paper consumption 65
potential 264 petrol use 60
power plant 250 sequestration 72
bio-oil 251 traffic 59
BtL fuel 253 carbon dioxide-free power plants 73

Renewable Energy and Climate Change Volker Quaschning


© 2010 John Wiley & Sons, Ltd

303
Index

CFC 36, 233 economics


chamber system 202 biomass 259
Chernoby l 10 geothermal energy 220
clean development mechanism 67 heat pump 232
climate change 24 hydrogen 272
climate protection 42 hydropower 206
coefficient of performance 225 photovoltaics 109
collector 118, 146 solar power plant 158
collector efficiency 119 solar thermal system 138
collector size 135 wind power 184
combined heat and power 75 efficiency
compression heat pump 226 biomass growth 239
concentrating photovoltaic 155 biomass power plant 249
concentration, solar radiation 157 collector 119
concentrator cell 155 fireplace 245
controlled ventilation 57 firewood boiler 246
convex mirror 146 fuel cell 272
open fireplace 245
dam 199 ORC power plant 217
damage from natural catastrophes 27 photovoltaics 90
deep drilling 213 steam turbine process 150
depth geothermal energy 210 electric car 284
depth gradient 213 electric stove 49
derrick 214 electricity generation costs
design hydro power 206
firewood boiler 256 photovoltaic 110
heat pump 228 solar power plant 163
hydropower plant 204 wind power 186
photovoltaics 103 electricity import 163
solar heating support 136 electrolysis 268
solar power plants 156 emissions trade 67
solar thermal system 133 energy consumption 13
solar-heating of drinking water 134 energy efficiency class 53
wind power 183 energy 2
wood pellet store 257 primary 14
direct-normal radiation intensity 157 reserves 20
Dish-Stirling power plant 153 saving 47
DNI 157 savings tips 54
domestic water heating 125
drag principle 171 fast breeder reactor 11
drilling fireplace 245
depth geometrics 213 firewood boiler 245
heating pump 231 fish bypass 207
Flatcon technology 155
Earth’s core 210 flat-plate collector 121
ecology float ball system 203
biomass 260 fossil energy sources 4
geothermal energy 220 Francis turbine 195
heat pump 233 frequency distribution 183
hydrogen production 273 Fresnel collector 146
hydropower 206 fuel cells 270
photovoltaics 112 fuel yield per hectare 262
solar power plant 160
solar thermal system 139 gas stove 49
wind power 187 geothermal depth gradient 213

304
Index

geothermal energy 210 storage power plant 198


ecology 220 tidal power plant 201
economics 220 turbines 194
HDR power plant 218 wave power plant 202
heat plant 215
heat pump 223 ice age 24
markets 221 ice coverage 27
planning 219 IPCC 37
power plant 216
global wind circulation 167 Jänschwalde lignite power plant 71
Goldisthal power plant 201 joint implementation 67
gravity solar system 125
green electricity 49
Kalina process 217
greenhouse effect 30
Kalkar power plant 11
greenhouse potential
Kaplan turbine 195
greenhouse gases 35
kites 289
refrigerants 234
Kvaerner process 267
Greenland ice 39
Kyoto protocol 44
gross domestic product 76
Kyrill storm 28
ground collector 229
Gulf Stream 41
laughing gas 35
Hadley cell 167 LH2, liquid hydrogen 269
Harrisburg 10 lift principle 171
HDR, hot dry rock 218 line concentrator 146
heat pipe 124 liquid hydrogen 269
heat pump 223 lithosphere 212
absorption 227 low-energy house 55
adsorption 228 low-energy light bulbs 52
coefficient of performance 225
compression 226 Manhattan project 9
ecology 233 markets
economics 232 biomass 263
heat sources 223 geothermal energy 221
planning 228 heat pump 235
markets 235 hydrogen 274
refrigerant 227, 234 hydropower 207
heat sources for heat pumps 224 photovoltaics 113
Helios solar aeroplane 290 solar power plants 161
HFC 34, 234 solar thermal system 140
hot dry rock power plant 218 wind power 188
Hurricane Katrina 28 maximum power point 91
hydrogen 266 methane 34
ecology 273 module price development 114
economics 272 molten carbonate fuel cell 272
markets 274 monocrystalline silicon 93
production 267 MPP, maximum power point 91
storage 269
hydropower 191 natural catastrophes 27
ecology 206 natural gas storage 8
economics 206 nuclear energy 9
markets 207 nuclear fusion 12
ocean current power plant 203
pumped-storage power plant 200 ocean current power plant 203
run-of-river power plant 197 offshore wind power 179

305
Index

oil power plant


barrel 6 biomass 248
crisis 22 brown coal 71
prices 23 carbon dioxide-free 73
reserves 20 combined heat and power 75
OPEC 5 combined renewable 85
open fireplace 245 concentrating photovoltaic 155
open-circuit voltage 91 Dish-Stirling 153
ORC power plant 216 geothermal 216
orientation hot dry rock (HDR) 218
photovoltaic system 174 hydro 197
oxygenic photosynthesis 238 Jänschwalde 71
oxyhydrogen reaction 265 Neurath 72
ozone 34 nuclear 12
ozone layer 35 ocean current 203
Organic Rankine Cycle (ORC) 216
parabolic trough power plant 147 parabolic trough 147
partial oxidation 267 photovoltaic 99
passive house 56 pumped-storage 200
Pelton turbine 196 run-of-river 197
PEMC, proton exchange membrane 271 SEGS 149
performance ratio 105 solar chimney 153
petrothermal geothermal energy 218 solar tower 150
photosynthesis 238 storage hydro 198
photovoltaic module 95 tidal 201
photovoltaics 87 wave 202
concentrator cell 155 wind 177
ecology 112 power-heat coupling 75
economics 109 primary energy 48
efficiency 90 primary energy requirements
functionality 88 biomass share 263
grid-coupled system 99 climate protection scenario 80
island grid system 96 development worldwide 13
markets 113 energy sources 16
module price development 114 heat supply 83
optimal orientation 104 per head 15
production 92 transport sector 84
planning 103 production
thin film 95 biodiesel 251
planning bioethanol 252
biomass heating 256 BtL fuel 253
firewood boiler 256 hydrogen 267
geothermal system 219 RME 251
heat pump 228 solar cell 92
hydropower plant 204 proton exchange membrane fuel cell 271
photovoltaic system 103 pumped-storage power plant 200
solar heating of drinking water 134 PV, see photovoltaics 87
solar heating support 136
solar power plant 156 receiver 150
solar thermal system 133 reduction targets 42
wind power 182 refrigerant 227, 234
plug-in hybrid car 284 refrigerating machine 227
plus-energy solar house 279 renewable energy import 163
point concentrator 147 Renewable Energy Law 69
polycrystalline solar cell 92 biomass power plants 260
power factor 171 geothermal power plants 220

306
Index

photovoltaics 111 solar radiation types 157


wind power 185 solar thermal system 116
renewable energy supply 84 design 133
RME 251 domestic water heating 125
rotor blade 172 ecology 139
round wood 242 economics 138
run-of-river power plants 197 heating support 128
markets 140
sailing ship 288 planning 133
sales of indulgences 67 power plants 144
sea level rise 38 solar tower power plant 150
secondary energy 48 solar-heated multiple-family house 279
consumption 50 solid cubic metre 243
traffic 59 solid oxide fuel cell 272
SEGS power plant 149 stacked cubic metre 243
selective absorber 122 standard test conditions 92
selective coating 121, 148 standby losses 51
semiconductor 88 STC, standard test conditions 92
short-circuit current 91 steam reforming 267
silicon 87 steam turbines 148
solar Stirling motor 153
absorber 120 storage hydro power plant 198
activity 29 storage size 135
car 285 stratosphere 36
cell 88 swimming pool absorber 120
chemistry 155 swimming pool heating 131
collector 121
cooker 133 tectonic plates 212
cooling 130 temperature change 34
district heating 130 temperatures, depths in Germany 213
domestic water heating 125 thermo-siphon solar system 126
energy 78 thin film photovoltaic module 174
energy import 163 three-litre house 56
ferry 289 tidal power plant 201
filling station 288 tilt gains 105
heating support 128 tower power plant 150
kitchen 292 tracking PV modules 100
mobile 285 trade wind 167
oven 145 traffic 59
plane 290 transport sector 58
share 134 trough power plant 147
swimming pool heating 131 turbine
solar power plant 144 bulb 195
concentrating photovoltaic 155 Francis 196
Dish-Stirling 153 Kaplan 194
ecology 160 Ossberger 197
economics 158 Pelton 196
markets 161 wind 177
parabolic-trough 147
photovoltaic 99
uranium supply 12
planning 156
useful energy 48
solar chimney 153
U-value 56
solar tower 150
solar radiation energy
Europe 104 vacuum flat-plate collector 123
worldwide 156 vacuum glazing 57

307
Index

vacuum insulation material 56 offshore 179


vacuum tube collector 123 park 178
ventilation system 57 planning 182
volumetric receiver 150 turbine 174
wind speed records 170
wafer 93 wind speed, velocity 169
warming period 24 window 56
water circulation on earth 192 wood 242
water turbine 194 wood briquettes 242
wave power plant 202 wood frame construction 56
wind global circulation 167 wood moisture 244
wind power 165 wood pellet heating 246
design 182 wood pellets 242
charger 173 prices 259
ecology 187 standard 241
economics 184 wood-burning stove 245
farm 178 world population 77
grid-coupled systems 174 World Solar Challenge 286
island system 174
lift principle 171 zero-emissions factory 283
markets 188 zero-heating cost house 280

308

You might also like